Download Activity Regulates the Synaptic Localization of the NMDA Receptor

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Biological neuron model wikipedia , lookup

Long-term potentiation wikipedia , lookup

Nervous system network models wikipedia , lookup

End-plate potential wikipedia , lookup

Axon guidance wikipedia , lookup

Neuroanatomy wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Development of the nervous system wikipedia , lookup

Apical dendrite wikipedia , lookup

Optogenetics wikipedia , lookup

Brain-derived neurotrophic factor wikipedia , lookup

Long-term depression wikipedia , lookup

Nonsynaptic plasticity wikipedia , lookup

Pre-Bötzinger complex wikipedia , lookup

Synaptic gating wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Neurotransmitter wikipedia , lookup

Signal transduction wikipedia , lookup

Neuromuscular junction wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Chemical synapse wikipedia , lookup

Synaptogenesis wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

NMDA receptor wikipedia , lookup

Transcript
Neuron, Vol. 19, 801–812, October, 1997, Copyright 1997 by Cell Press
Activity Regulates the Synaptic Localization
of the NMDA Receptor in Hippocampal Neurons
Anuradha Rao and Ann Marie Craig*
Department of Cell and Structural Biology
University of Illinois
Urbana, Illinois 61801
Summary
We describe here a novel effect of activity on the subcellular distribution of NMDA receptors in hippocampal neurons in culture. In spontaneously active neurons,
NMDA receptors were clustered at a few synaptic and
nonsynaptic sites. Chronic blockade of NMDA receptor activity induced a 380% increase in the number of
NMDA receptor clusters and a shift to a more synaptic
distribution. This effect was reversible. The distributions of the presynaptic marker synaptophysin, the
AMPA-type glutamate receptor subunit GluR1, and the
putative NMDA receptor clustering protein PSD-95
were not affected by blockade. Regulation of the synaptic localization of NMDA receptors by activity may
define a novel mechanism by which input controls a
neuron’s ability to modify its synapses.
Introduction
The N-methyl-D-aspartate (NMDA)-type glutamate receptors comprise a class of ionotropic glutamate receptors that play a central role in synaptic plasticity, circuit
development, memory formation, and excitotoxicity (Bliss
and Collingridge, 1993; Choi, 1994; Scheetz and Constantine-Paton, 1994; Tsien et al., 1996). Molecular cloning has revealed that NMDA receptors are heteromeric
ion channels composed of the essential NR1 subunit
(Moriyoshi et al., 1991; Forrest et al., 1994; Li et al.,
1994), which exists in several alternatively spliced forms
(Durand et al., 1993; Laurie and Seeburg, 1994), and
one or more of the modulatory NR2 subunits, NR2A–D
(Kutsuwada et al., 1992; Monyer et al., 1992, 1994; Ishii
et al., 1993). Regional and developmental differences in
heteromeric composition are believed to change functional properties of the receptor, such as agonist affinities,
Mg2 1 sensitivity, Ca 21 permeability, channel conductance,
deactivation kinetics, and response to modulatory agents
(reviewed by McBain and Mayer, 1994; Feldmeyer and
Cull-Candy, 1996).
The properties of the NMDA receptor channel can be
rapidly regulated by activity during synaptic transmission. The NMDA responsiveness and NMDA receptor
density of neurons can also be modulated over a time
course of days to weeks by activity, both in vivo and in
primary neuron cultures. These slow changes in NMDA
responsiveness may be developmentally restricted. In
the visual cortex, the NMDA responsiveness of neurons
is high during the critical period for ocular dominance
column formation and decreases greatly at the end of
the critical period (Fox et al., 1991; Carmignato and
* To whom correspondence should be addressed.
Vicini, 1992). Activity blockade prevents the decrease
in NMDA responsiveness and extends the critical period
for synaptic reorganization. Conversely, frog tecta chronically treated with NMDA show a decreased NMDA responsiveness coupled to refinement of topographic
map formation (Debski et al., 1991).
Changes in NMDA responsiveness in cultured neurons have been linked to changes in levels of different
subunits of the receptor. In cultured cerebellar granule
cells, calcium influx via potassium or NMDA-induced
depolarization is necessary for cell survival early in development and elicits an increase in NMDA responsiveness. This is associated with a change in subunit
composition of the NMDA receptor, particularly an increase in NR2A and a decrease in NR2B subunits (Bessho et al., 1994; Vallano et al., 1996). Later in development, NMDA responsiveness and NMDA receptor density
can be down-regulated by NMDA treatment (Didier et al.,
1994; Resink et al., 1995). In cultured cortical neurons,
NMDA receptor density and specifically NR2B mRNA
and protein levels can be up-regulated by treatment
with NMDA receptor antagonists (Williams et al., 1992;
Follesa and Ticku, 1996).
Electrophysiological studies of neurons in culture suggest that the NMDA receptor is present in synaptic “hot
spots” on dendrites and that NMDA receptors and nonNMDA receptors are frequently but not always colocalized at such synaptic sites (Bekkers and Stevens, 1989;
Jones and Baughman, 1991). Immunocytochemical studies at both light and electron microscopic levels have
demonstrated NMDA receptor immunoreactivity at the
postsynaptic membrane of spiny excitatory sites in adult
brain (Petralia et al., 1994a, 1994b; Siegel et al., 1994;
Johnson et al., 1996) as well as at extrasynaptic plasma
membrane in neonatal cortex (Aoki et al., 1994). Doublelabel immunolocalization with antibodies against NR1 and
AMPA-type glutamate receptors indicate that the two receptors are colocalized at the postsynaptic membrane of
some synapses (Siegel et al., 1995; Kharazia et al., 1996).
We set out to determine whether the subcellular distribution of NMDA receptors can be modulated by activity.
In cultured hippocampal neurons, we found that chronic
activity blockade changes the subcellular distribution
of the NMDA receptor with no effect on the AMPA-type
glutamate receptor. Activity blockade induced an increase both in the total number of NMDA receptor clusters (hot spots) and in their targeting to synaptic sites.
Results
NMDA Receptors Exhibit Different Patterns
of Localization in Hippocampal
Pyramidal Neurons
Embryonic rat hippocampal pyramidal cells grown in low
density cultures were used to examine NMDA receptor
distribution. At 3 weeks in culture, over 90% of the neurons showed nonuniform immunoreactivity for the NR1
subunit. Bright clusters of NR1 staining were present,
typically on dendritic shafts of fine branches at the periphery of the dendritic tree (Figure 1A). Comparison
Neuron
802
Figure 1. Treatment of Hippocampal Neurons with the NMDA Receptor Antagonist APV Changes the Distribution of NMDAR1
(A and B) Control cells; (C and D) APV-treated cells; (E and F) reversal. Neurons were treated with control media (A and B) or APV (C–F) during
weeks 2–3 in culture and then fixed at the end of week 3 (A–D) or placed back in control media for a fourth week (E and F) and then
immunolabeled for NR1. The control cells in (A) and (B) show the distal and the proximal dendritic shaft NR1 clustering patterns, respectively.
The APV-treated cells in (C) and (D) display the ubiquitous spiny clusters seen in most cells after APV treatment and very seldom without it.
After a fourth week allowing spontaneous activity to resume, the neurons again mostly displayed the distal dendrite shaft NR1 clustering
pattern (E and F). Scale bar, 10 mm.
of NR1 immunoreactivity to the distribution of synaptic
vesicle clusters, revealed by immunostaining for synaptophysin (Fletcher et al., 1991), indicated that these distal dendritic clusters were usually synaptic. Two other
patterns of NR1 immunoreactivity were observed less
frequently. On some cells, bright clusters of NR1 were
arrayed in the cell body and proximal dendrite shafts
(Figure 1B). These clusters were generally nonsynaptic
(Figures 2A–2D) and were more prominent in immature
neurons. Similar patches of NR1 immunoreactivity have
been observed in vivo in the core of apical dendrites
(Johnson et al., 1996) and at the plasma membrane of
soma and dendrites of developing neurons (Aoki et al.,
1994). The third pattern, seen rarely at 3 weeks but more
frequently at 5–6 weeks, was of numerous synaptic and
often spiny clusters along the length of dendrites. From
analyses of these cultures at different ages (A. R. and
A. M. C., unpublished data), we found a developmental
progression in the distribution of NR1 from cell body
and proximal dendrite nonsynaptic clusters to synaptic
clusters on distal dendrite shafts to spiny synaptic clusters throughout the dendritic length.
NMDA Receptor Blockade Increases NMDA
Receptor Cluster Number and
Synaptic Localization
Chronic treatment of hippocampal cultures with the
NMDA receptor antagonist 2-amino-5-phosphonovalerate (APV) caused a dramatic shift in the pattern of NR1
immunoreactivity such that most cells now had numerous synaptic and often spiny clusters of NR1, even at
3 weeks in culture (Figures 1C, 1D, 2E–2H). This effect
was specific to pyramidal neurons and not GABAergic
neurons. We compared the number and location of NR1
clusters in sets of randomly selected control and APVtreated pyramidal neurons (Figure 3). There was a 380%
increase in the number of NMDA receptor clusters from
9.8 6 1.3 clusters/100 mm dendrite length in control
cells to 37.5 6 2.6 clusters/100 mm in APV-treated cells
(mean 6 SEM, n 5 39, significant at p , 0.001). This
effect of APV was due solely to an increase in the number
of NR1 clusters at synaptic sites and was accompanied
by a slight decrease in the number of NR1 clusters at
nonsynaptic sites. We saw no increase in the intensity
of NR1 staining at terminal shaft clusters in the APVtreated cells, and bright nonsynaptic arrays of clusters
in the cell body and proximal dendrites were no longer
evident, implying that there is not a generalized increase
in the amount of NR1 at all sites but indeed a shift in
the distribution. There was no change in the pattern
of synaptophysin immunoreactivity or in the number of
synaptophysin-labeled presynaptic terminals after APV
treatment (Figures 2, 5, and 6). The conclusion from
these results is that APV induces an increase in NR1 at
previously existing synaptic sites. The observed change
in NR1 localization could involve relocation of previously
existing receptor molecules from nonsynaptic to synaptic
sites, or a shift in targeting of newly synthesized receptor.
Activity Effect on NMDA Receptor Clustering
803
Figure 2. APV Treatment Increases NMDA
Receptor Clustering at Synaptic Sites
Double-label immunocytochemistry for NR1
(green) and synaptophysin (red); yellow in the
superimposed images (A, D, E, and H) indicates colocalization. In the control cells, the
proximal dendrite NR1 clusters do not colocalize with synaptophysin clusters (A–D). In
the APV-treated cells, the ubiquitous spiny
NR1 clusters are mostly at synaptophysinlabeled synapses (E–H). Scale bar, 10 mm (A
and E); (B)–(D) and (F)–(H) show enlarged
views of single dendrites from (A) and (E),
respectively.
Surprisingly, Western blot analysis (Figure 7) showed
no change in the amount of NR1 in the APV-treated
compared to control cultures. Comparing the total intensity/cluster in control proximal or distal dendrite shaft
clusters versus APV-treated spiny clusters, we found
that control dendrite shaft clusters (n 5 968) had on
average 254% greater total intensity than did APV spiny
clusters (n 5 2214). However, summing the total fluorescent intensity in all clusters in the two conditions (both
shaft and spine clusters corresponding to data shown
in Figure 3), we observed that total NR1 intensity in
control cells was still only 44% of that in APV-treated
cells. Where is the other 56% of the NR1 protein in the
control cells if it is not at clusters? Hall and Soderling
(1997) showed that whereas all of the NR2B is at the
cell surface, only 50% of NR1 is exposed to the cell
surface in cultured hippocampal neurons. It is possible
that some of the NR1 in our control neurons is in a
diffuse intracellular pool, which in APV-treated cells becomes part of the synaptic pool. Levels of diffuse staining for NR1 were low and could not be reliably measured
for differences between conditions.
The induction of synaptic NR1 clusters by NMDA receptor blockade implies that spontaneous activity in the
cultures normally drives the NMDA receptors away from
synapses. To test this directly, we incubated 3 week
APV-treated neurons (with synaptic NMDA receptor
clusters) for a fourth week in the absence of APV, thus
Figure 3. NMDA Receptor Blockade Increases NMDA Receptor Cluster Number and
Synaptic Localization
NR1 clusters were counted in matched controls and neurons treated with APV, CNQX,
nifedipine, TTX, and TTX plus NMDA. Each
NR1 cluster was classified as synaptic or nonsynaptic based on the presence or absence
of synaptophysin immunoreactivity. Two dendrites in each of ten randomly selected cells
were evaluated in each of three separate experiments per treatment. The asterisks indicate treatments that resulted in significant
differences in cluster number when compared to the matched control (t test, p , 0.01,
except nifedipine, where p 5 0.034). The effect of TTX plus NMDA treatment was also
significantly different from the TTX treatment
(p , 0.01).
Neuron
804
Figure 4. NMDA Receptor Blockade Has No Effect on the Distribution of the AMPA-Type Glutamate Receptor Subunit GluR1
In cells fixed in paraformaldehyde, GluR1 clusters were well-preserved and showed a similar pattern of immunoreactivity in control
cells (A) versus APV-treated cells (B). (C) GluR1 clusters were
counted in matched controls and neurons treated with APV or TTX.
Two dendrites in each of ten randomly selected cells were evaluated
in each of two separate experiments per treatment. There was no
significant difference in the number of GluR1 clusters. Scale bar,
10 mm.
allowing the cells to return to their normal levels of spontaneous activity. The resumption of spontaneous activity
resulted in a large decrease in synaptic NR1 clusters
compared to the 3- or 4 week APV-treated neurons (Figures 1E and 1F). In fact, the number of NR1 clusters now
appeared to be lower than in a 4 week untreated culture,
presumably due to the enhanced activity through the
increased number of synaptic NMDA receptor clusters
resulting from the initial APV treatment. These results
indicate that stimulation can also induce a redistribution
of NMDA receptors away from existing synaptic sites.
Activity of the NMDA receptor could mediate the redistribution of NMDAR1 through an effect on postsynaptic depolarization, intracellular calcium levels, localized
calcium entry, or NMDA receptor conformation. We reduced postsynaptic depolarization by treating cultures
with the voltage-dependent sodium channel blocker tetrodotoxin (TTX) or the AMPA-type glutamate receptor
antagonist 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX)
and assessed the patterns of NR1 and synaptophysin
immunostaining (Figure 3). Chronic treatment with TTX
had a similar effect to that of APV, causing a 320%
increase in NR1 cluster number and a shift toward more
synaptic clusters. CNQX caused a smaller increase in
NR1 cluster number and shift to synaptic sites. While
TTX and CNQX reduce postsynaptic depolarization, they
also indirectly reduce NMDA receptor activation. Addition of 5 mM NMDA largely blocked the increase in NR1
cluster number and shift to synaptic sites induced by
TTX, suggesting that the effect of TTX was primarily
due to blockade of NMDA receptor activation and not
postsynaptic depolarization. One of the major effects of
NMDA receptor activation is localized calcium entry. In
addition to NMDA receptors, L-type voltage-dependent
calcium channels are the other major pathway for calcium entry into the somatodendritic domain of hippocampal neurons, but calcium entry through these two
pathways activates different signaling cascades (Bading
et al., 1993). K1-induced calcium accumulation in the
somata and dendritic spines of mature hippocampal
Figure 5. NMDA Receptor Blockade Results in an Increase in the
Number of Synapses with Clustered NR1 and GluR1 as Well as in
Synapses Containing NR1 Alone (Potential Silent Synapses)
Neurons were fixed in methanol and triple-stained with antibodies
to synaptophysin (blue), NR1 (red), and GluR1 (green). In control
cells (A–D), NR1 (A) and GluR1 (B) clustered at distinct synapses,
NR1 on distal dendritic shafts, and GluR1 on spines (see superimposed image in [C]). In APV-treated cells (E–H), NR1 clustered at
GluR1-containing spines (yellow in superimposed image in [G]) as
well as at additional spine synapses (red in [G]). Scale bar, 10 mm.
neurons can be greatly reduced by nifedipine, an L-type
voltage-dependent calcium channel antagonist (Basarsky et al., 1994; Segal, 1995). We found that chronic
treatment with nifedipine did not significantly affect the
number of synaptic NR1 clusters (Figure 3). Taken together, these results suggest that it is specifically the
level of activity of the NMDA receptor that regulates
the subcellular distribution of NR1. This effect may be
mediated by localized calcium entry through the NMDA
receptor channel or by a direct effect on NMDA receptor
conformation and thus its interaction with other proteins.
Activity Blockade Has No Effect on the Distribution
of AMPA-Type Glutamate Receptors
The effect of NMDA receptor blockade was specific
for NR1 when compared to another type of glutamate
receptor, the GluR1 subunit of the AMPA-type glutamate
receptor. GluR1 and GluR2/3 subunits completely colocalize at individual synapses in these neurons, and
GluR4 is not expressed at high levels, so GluR1 can be
considered a marker of AMPA receptor synapses (Craig
et al., 1993). There was no difference in the number
of GluR1 clusters between control cultures and those
treated with APV (Figures 4A–4C and 6) or with TTX
(Figure 4C).
We compared the localizations of GluR1, NR1, and
synaptophysin in control and APV-treated pyramidal
neurons (Figures 5 and 6). GluR1 clusters were always
Activity Effect on NMDA Receptor Clustering
805
it is important to note that all of the synapses classed
as NR1 GR2 lacked GluR1 clusters. Although most had
no detectable immunoreactivity for GluR1, a few had
faint but detectable immunoreactivity for GluR1 in the
spine at the same intensity levels as the diffuse staining
in the dendrite shaft, which could not be classified as
clustered staining. This diffuse level of staining may indicate the presence of AMPA receptor at these synapses,
making them functional rather than silent. The relative
strength of these synapses is likely to be different.
Figure 6. NMDA Receptor Blockade Changes the Glutamate Receptor Composition of Synapses, but Not the Total Number of Synapses Defined by Presynaptic Specializations
The number of synaptophysin (synp), GluR1, and NR1 clusters per
length of dendrite was measured for triple-labeled pyramidal cells
from matched APV-treated and control cultures. For each GluR1 or
NR1 cluster observed, it was classified as containing hotspots of
immunoreactivity for both GluR1 and NR1 (GR1 NR1), for GluR1
but not NR1 (GR1 NR-), or for NR1 but not GluR1 (GR2 NR1). All
of the GluR1 clusters were synaptic, whereas only 78% of control
and 96% of APV-treated NR1 clusters were synaptic (see Figure 3).
Two dendrites in each of ten randomly selected cells were evaluated
in each of two separate experiments per treatment. The asterisks
indicate significant differences in cluster number in APV-treated
cells when compared to the matched control (t test, p , 0.001); for
synaptophysin and GluR1, there was no significant difference (t test,
p . 0.01).
synaptic and did not change distribution with APV treatment (compare Figure 5B with Figure 5F), but the relation
between GluR1 and NR1 clusters changed due to the
shift in NR1 localization (compare Figures 5A and 5C
with Figures 5E and 5G). In control spontaneously active
neurons, we calculated (by combining the data in Figures 3 and 6) that 49% of the synaptic sites defined
by synaptophysin clusters were not apposed to either
GluR1 or NR1 clusters, 23% were apposed to GluR1
clusters alone, 19% to NR1 clusters alone, and 9% were
apposed to postsynaptic sites containing clusters of
both GluR1 and NR1. The increase in NR1 clustering at
synaptic sites after NMDA receptor blockade resulted
in a large increase in the number of synapses containing
concentrations of NR1 alone (red spots in Figure 5G;
compare with the control cell in Figure 5C), such that
they comprised 62% of total synapses (Figure 6). The
APV-induced increase in NR1 clustering also resulted
in a doubling of synapses containing concentrations of
both NR1 and GluR1 (yellow spots in Figure 5G) to 20%
of the total synapses and a decrease in synapses containing clusters of GluR1 but not NR1 to one-fifth of the
control number (4% of total synapses). Thus, after APV
treatment, NR1 clusters are at 82% of the total synapses
and probably at almost all the glutamatergic synapses.
The synapses containing clusters of NR1 without concentrations of GluR1 that we see in control cells and in
much greater numbers in APV-treated cells may provide
a morphological basis for the “silent synapses” suggested by physiological studies (Isaac et al., 1995; Kullmann and Siegelbaum, 1995; Liao et al., 1995). However,
Potential Mechanisms of the APV-Induced
Redistribution of NMDA Receptors: Roles
of NR2 Subunits and PSD-95
Potential molecular mechanisms for the regulation of
NMDA receptor distribution include a switch in expression of a splice variant form of NR1 (which has been
shown to regulate localization of NR1 in transfected
fibroblasts; Ehlers et al., 1995) or NR2 subunit, the accumulation of a posttranslational modification such as phosphorylation, and/or a change in the interaction of the
NMDA receptor with associated proteins such as PSD-95
(Kornau et al., 1995) and a-actinin (Wyszynski et al.,
1997). One clue to the mechanisms of NR1 redistribution
was the relatively slow time course: a 3 day APV treatment induced an almost complete redistribution, a 1
day APV treatment had only a partial effect, and a 4 hr
treatment showed no detectable effect (data not shown).
This time course suggests that APV acts by inducing
transcriptional or translational changes rather than
faster posttranslational changes.
We determined whether APV induces any change in
the level of expression of NR1 (all splice forms), NR1
containing the C1 exon (NR1-C1), NR1 containing the
N1 exon (NR1-N1), and the two NR2 subunits that are
highly expressed in the hippocampus (NR2A and NR2B;
Monyer et al., 1994; Petralia et al., 1994b) by Western
blotting of culture extracts (Figure 7). There was no difference in the amounts of NR1 or the two splice variant
forms of NR1 between control and APV-treated cultures,
but there was a .200% increase in the amounts of both
NR2A and NR2B after APV treatment.
The correlation between higher NR2 subunit levels
and a more synaptic receptor distribution after APV
treatment suggested that a shift to a higher ratio of NR2type subunits in the NMDA receptor oligomer may be
responsible for the redistribution of the complex to synaptic sites. To test this more directly, we double labeled
neurons with antibodies to NR1 and either NR2A or
NR2B (Figure 8). In APV-treated cells, NR2A and NR2B
clusters increased dramatically in number, and both proteins were colocalized with NR1 at synaptic sites. However, both proteins were also present with NR1 in the
proximal and distal dendrite shaft clusters in control
cells (many of which are nonsynaptic, as shown in Figure
2). We determined the fluorescence intensity ratios of
the NR2A or NR2B to NR1 subunits at the primarily
nonsynaptic control cell dendrite shaft clusters versus
the spiny and primarily synaptic APV-treated cell clusters (Table 1). There was no detectable difference in the
ratio of the NR2 to NR1 subunits in clusters between
the control and APV groups. The observation that NR2A
Neuron
806
not shown). Thus, NMDA receptor blockade does not
affect the distribution of PSD-95 but does induce a redistribution of NMDA receptors to synaptic sites previously
containing PSD-95. These data indicate that the APVinduced redistribution of NMDA receptors is not mediated by a change in localization of PSD-95 but are still
consistent with the idea that PSD-95 may be necessary
for synaptic localization of NMDA receptors. The differential localization of PSD-95 and NMDA receptors in
control cells indicates that the presence of PSD-95 at
synapses is not always sufficient to induce synaptic anchoring of NMDA receptors. a-Actinin-2, an actin-binding
protein that binds to NR1 and NR2B and is concentrated
in dendritic spines (Wyszynski et al., 1997), also showed
no change in distribution after APV treatment (data not
shown). Thus, like PSD-95, a-actinin-2 is not sufficient
but may be important for anchoring the NMDA receptor
at synapses in hippocampal pyramidal neurons.
Discussion
Figure 7. The APV-Induced Redistribution of NR1 Is Correlated with
Increased Levels of NR2 Subunits in Western Blots of Control Compared to APV-Treated Hippocampal Cultures
Control (C) and APV-treated (A) lanes show no difference in signal
intensity for total NR1 or two NR1 splice variants, but do show an
increase in signal for NR2A and NR2B subunits with APV treatment.
The graph shows the average of relative signal intensity values from
Western blots of cell homogenates from three different experiments
normalized with respect to tubulin signals from the same lane; values
reported here are for APV treatment as a percentage of control for
the five antibodies used.
or NR2B and NR1 have the same stoichiometry in synaptic and nonsynaptic clusters indicates that a change in
stoichiometry within the oligomer is not responsible for
the new pattern of localization after APV treatment.
Thus, the increase in NR2 levels cannot be solely responsible for the redistribution of the receptor complex.
The change in levels of NR2 may still contribute to synaptic localization of the complex by increasing surface
accessibility and/or by mediating binding to potential
synaptic anchoring proteins.
The NR2 subunits bind to PSD-95, a PDZ domain
protein highly concentrated in the postsynaptic densities of neurons (Cho et al., 1992; Kornau et al., 1995).
PSD-95 coclusters with NR2 subunits upon coexpression in nonneuronal cells, but neither protein clusters
when expressed alone (Kim et al., 1996). PSD-95 has
been suggested to function as a synaptic NMDA receptor clustering protein, but its function in neurons has
not yet been established.
To test the role of PSD-95 in the synaptic localization
of NMDA receptors, we compared the pattern of PSD95 immunoreactivity with that of NR1 in control and APVtreated neurons (Figure 9). There were a large number
of PSD-95 clusters in control cells, and most of these
did not display immunoreactivity for NR1 (Figure 9A; for
NR2A or NR2B, data not shown). There was no change
in the number of PSD-95 clusters in APV-treated cells,
but in this case, the PSD-95 clusters colocalized with
the NR1 clusters (Figure 9B; with NR2A and NR2B, data
In summary, we describe here a novel activity-dependent form of synaptic plasticity at the molecular and
cellular level in an accessible culture system. We found
that the subcellular distribution of the NMDA receptor in
hippocampal pyramidal neurons is regulated by activity,
such that activity blockade induces an increase in receptor clustering at synaptic sites, while spontaneous
activity results in a decrease in synaptic clustering. By
pharmacological manipulation, we found that NMDA receptor localization was largely controlled by the activity
of the NMDA receptor itself, perhaps through localized
changes in calcium entry or direct changes in NMDA
receptor conformation. Activity blockade did not affect
the total number of synapses or the subcellular distribution of two other excitatory postsynaptic proteins, the
AMPA-type glutamate receptor subunit GluR1 and the
PDZ domain protein PSD-95. Activity selectively regulated the number of NMDA receptor clusters at existing
synaptic sites. Furthermore, we found that the distributions of NMDA and AMPA-type glutamate receptors are
independently regulated and that these two receptor
types can cluster at different postsynaptic sites on a
single neuron. The APV-induced synaptic NMDA receptor pattern correlated with a selective increase in overall
levels of the NR2A and NR2B subunits. However, the
NR2:NR1 ratio in the oligomer was not different at APVinduced synaptic sites compared with control nonsynaptic sites, indicating that additional mechanisms regulate
the synaptic redistribution of the complex. Finally, we
found that the presence of the NR2-binding protein PSD95 at synapses may be necessary but is not sufficient to
induce synaptic localization of the NMDA receptor.
These results provide evidence for a regulated subcellular distribution of the NMDA receptor in neurons.
Short-term posttranslational modifications and increases
or decreases in expression levels appear to be common
methods by which neurons regulate their ion channel
complements. We describe here a novel mechanism of
feedback regulation of subcellular distribution of an ion
channel by its own activity. Given the essential role of
the NMDA receptor in many forms of synaptic plasticity,
Activity Effect on NMDA Receptor Clustering
807
Figure 8. APV Treatment Increases the Number of NR2A and NR2B Clusters in Parallel
with NR1
Double staining is shown here for NR1 (red)
and NR2A or NR2B (green) in control (left)
versus APV-treated (right) neurons. Regions
of the red and green images corresponding to
each superimposed image are shown slightly
enlarged at the bottom of each of the panels.
NR2A (A and B) and NR2B (C and D) are colocalized with NR1 at proximal dendrite clusters
in control cells (A and C) as well as at spiny
synaptic clusters in APV-treated cells (B and
D). Scale bar, 10 mm.
the activity-dependent regulation of synaptic localization of NMDA receptors may be a means of regulating
the capacity of a neuron to undergo further activitydependent synaptic modification.
Relationship between NMDA Receptors
and AMPA Receptors
It is not known whether all types of glutamate receptors
are colocalized within a neuron or whether there may
be differential targeting to individual postsynaptic sites.
Light and electron microscopic analyses of the distributions of different classes of glutamate receptors in brain
sections have revealed a colocalization of NMDA and
AMPA-type receptors at some spiny postsynaptic sites
(Siegel et al., 1995; Kharazia et al., 1996). While it has
been suggested that NMDAR1 may be localized at only
a subset of glutamatergic synaptic sites (Huntley et al.,
1994; Petralia et al., 1994a), these kinds of tissue localization studies cannot easily show the degree of overlap
(or lack thereof) of the two receptor types throughout
individual neurons. We were able to compare the overall
distributions of AMPA and NMDA receptors using cultured neurons. Surprisingly, these two receptor types
were largely not colocalized in control spontaneously
active cultures. Most of the AMPA receptor clusters on
dendritic spines lacked concentrations of NMDA receptors. Twenty-two percent of the NMDA receptor clusters
were nonsynaptic, and these lacked concentrations of
AMPA receptors. A substantial number of synaptic NMDA
receptor clusters (67%), particularly those on distal dendrite shafts, also lacked clusters of AMPA receptors
(Figures 5A–5D and 6). Thus, even the synaptic populations of NMDA and AMPA receptors in control neurons
were largely nonoverlapping. The observation of synaptic NMDA receptor clusters lacking detectable levels of
AMPA receptors also constitutes evidence of a morphological substrate for the silent synapses suggested by
physiological studies (Isaac et al., 1995; Kullmann and
Siegelbaum, 1995; Liao et al., 1995). The number of
potential silent synapses containing clusters of NMDA
but not AMPA receptors increased with chronic APV
treatment as did the number of synapses containing
clusters of both NMDA and AMPA receptors (Figures
5E–5H and 6). Thus, both in spontaneously active neurons and under conditions of activity blockade, NMDA
and AMPA receptor distributions were distinct, indicating that different mechanisms regulate their subcellular
targeting.
Mechanisms of the Activity Effect on NMDA
Receptor Distribution
The effect of activity blockade on NMDA receptor distribution correlated with an increase in the amount of two
Table 1. Ratio of NR2/NR1 Fluorescence Intensity in Clusters
Intensity Ratio
NR2B/NR1
Mean 6 SD
n, clusters (cells)
NR2A/NR1
Mean 6 SD
n, clusters (cells)
Control
APV
1.00 6 0.47
590 (32)
0.95 6 0.33
1396 (15)
1.00 6 0.52
380 (20)
1.03 6 0.32
818 (10)
The NR2B:NR1 and NR2A:NR1 fluorescence intensity ratios were
measured in sets of matched control versus APV-treated neurons.
Cells were chosen by scanning in the NR1 channel for typical proximal or distal control clusters or for typical spiny APV clusters and
then measured for the subunit fluorescence intensity ratios; values
reported here are normalized to the control groups. There was no
difference in subunit ratios between control and APV groups (t test,
p . 0.01).
Neuron
808
Figure 9. APV Treatment Does Not Affect the
Distribution of PSD-95
Double staining is shown here for NR1 (red)
and PSD-95 (green) in control (A) versus APVtreated (B) neurons. Regions of the red and
green images corresponding to each superimposed image are shown slightly enlarged
at the bottom of each of the panels. PSD-95
distribution does not change with APV treatment. PSD-95 is clustered at many sites in
control cells (A) and most of these clusters
lack NR1. APV treatment (B) results in an increase in NR1 clusters, which now colocalize
with PSD-95. Scale bar, 10 mm.
subunits of the NMDA receptor, NR2A and NR2B, with
no change in the level of the NR1 subunit. One hypothesis is that changes in levels of Ca2 1 influx through the
NMDA receptor initiate signal transduction cascades
that regulate NR2 transcription (among other events)
and thus mediate the redistribution of NMDA receptors.
There are many precedents for regulation of gene expression by Ca21 entry through the NMDA receptor
channel (Bading et al., 1993; Ghosh and Greenberg,
1995). Furthermore, NR2 mRNA levels are regulated by
activity in cortical and cerebellar granule neurons (Bessho et al., 1994; Follesa and Ticku, 1996; Vallano et al.,
1996). However, the finding that NR2A and NR2B were
both present at the same ratios in control nonsynaptic
NR1 clusters as in APV-induced synaptic NMDA receptor clusters indicates that the enhanced levels of NR2
subunit are not sufficient to mediate the change in synaptic localization of the complex. The enhanced levels
of NR2 may still be one of the factors contributing to
the synaptic localization, perhaps by simply increasing
the total surface pool of receptors. NR2 subunits can
increase the cell surface expression of NR1 in transfected nonneuronal cells (McIlhinney et al., 1996) and
show a greater degree of cell surface localization than
NR1 in cultured hippocampal neurons (Hall and Soderling, 1997). NR2 subunits may also contribute to the
synaptic localization of NMDA receptors by virtue of
their binding to PSD-95.
PSD-95 is a member of the PDZ family of proteins,
which are believed to be important in the formation of
cell junctional specializations (reviewed in Gomperts,
1996). PSD-95/SAP90 was originally isolated due to its
presence in postsynaptic density or synaptic junction
subcellular fractions (Cho et al., 1992; Kistner et al.,
1993). It was subsequently found to interact with the
carboxy-terminal tails of a K1 channel (Kim et al., 1995)
and of the NR2A and NR2B subunits of the NMDA receptor (Kornau et al., 1995; Niethammer et al., 1996).
PSD-95 colocalizes with NR2B at synaptic sites in cultured hippocampal neurons (Kornau et al., 1995) and
coimmunoprecipitates with NR2B from rat brain (Wyszynski et al., 1997). Coexpression of PSD-95 along with
either the K1 channel KV1.4 or NR2A/B in heterologous
cells is sufficient to induce the formation of clusters
containing both proteins (Kim et al., 1995, 1996). Thus,
PSD-95 is an attractive candidate for clustering NMDA
receptors in neurons.
We showed here that PSD-95 can form clusters at
synaptic sites that do not show concentrations of NMDA
receptor immunoreactivity (in control spontaneously
active cultures; Figure 9A). This result indicates that
synaptic clustering of PSD-95 is not sufficient to induce
synaptic anchoring of NMDA receptors. After APV treatment, NMDA receptors were coclustered with PSD-95
at synapses (Figure 9B), consistent with the idea that
PSD-95 may be necessary for synaptic localization of
the NMDA receptor.
The mechanism of the APV effect cannot be a simple
increase in NR2 subunits leading to increased PSD-95
binding, since NR2 subunits were present in equal ratios
at the nonsynaptic and synaptic NMDA receptor clusters. An additional step must be necessary to mediate
the synaptic localization of the complex. One interesting
possibility is that posttranslational modifications may
regulate the interaction between NR2 subunits and
PSD-95 family members. Cohen et al. (1996) found that
binding of PSD-95 to the inward rectifier K1 channel Kir
2.3 is regulated by protein kinase A phosphorylation of
a serine within the Kir 2.3 C-terminal PDZ domain binding site. Although NR2 subunits lack this consensus
phosphorylation site at their C terminus, other phosphorylation events modifying NR2 or PSD-95 or novel
anchoring proteins may be regulatory. Given the prolonged time course required for this effect, an attractive
hypothesis would be that APV induces an increase in the
synthesis of a regulatory enzyme that posttranslationally
modifies the receptor or its synaptic binding partner. In
this scenario, it is possible that the increase in NR2
levels is more of a consequence than a cause of the
redistribution; for example, NR2 may be stabilized by
buildup of a synaptic posttranslational modification of
NR2 or interacting proteins. Additional analyses of the
synthesis, stability, and posttranslational changes in
NR2 subunits and interacting proteins will be required to
elucidate fully the mechanisms involved in APV-induced
redistribution of the NMDA receptor.
Comparison with Activity-Driven Effects on AChR
Clustering at the Neuromuscular Junction
A well-studied example of activity-dependent regulation
of neurotransmitter receptor distribution is at the vertebrate neuromuscular junction (NMJ). Activity blockade
of innervated muscle results in an increase in transcription of acetylcholine receptor (AChR) subunits by nonsynaptic nuclei, thus leading to an increase in the
Activity Effect on NMDA Receptor Clustering
809
amount of extrajunctional receptor (reviewed by Hall
and Sanes, 1993; Duclert and Changeaux, 1995). The
effect of activity blockade can be reversed by direct
stimulation of the muscle. Activity-dependent regulatory
elements have been found in the promoters of AChR
genes. In this study, we similarly found an increase in
the total number of NMDA receptor clusters after activity
blockade of neurons, coupled to an increase in synthesis
of some subunits of the receptor. However, in contrast
to the increase in extrasynaptic AChR seen in muscle,
activity blockade induced a selective increase in synaptic NMDA receptors in neurons. This difference may be
due to the contrasting mechanisms of compartmentalization, by selective transcription in the multinucleate
myotubes versus by selective protein targeting in the
neurons.
A more subtle effect of activity on AChR clustering
was found in studies of developmental synapse elimination at the NMJ using focal postsynaptic blockade induced by local application of a-bungarotoxin (BaliceGordon and Lichtman, 1994). Local activity blockade in
small regions of the junction induced a gradual loss of
receptors from the blocked regions, followed by removal
of the overlying nerve terminal. However, blockade of
the entire junction or large parts of it resulted in no
change. In addition, activity blockade results in retention
of polyneuronal innervation of skeletal muscle, whereas
activity facilitates synapse elimination to a singly innervated state (reviewed in Nguyen and Lichtman, 1996).
Based on these findings, a model of activity-dependent
synapse elimination was proposed, where asynchronous activity of competing terminals at a postsynaptic
surface could result in an active terminal generating a
diffuse “punishment” signal that induces destabilization
of adjacent synapses in the absence of a localized protective signal also generated by activity (Nguyen and
Lichtman, 1996). The effect of activity blockade on
NMDA receptor distribution may be similar to that in the
NMJ, in that the postsynaptic receptors are stable in
the absence of activity in both cases. Further work will
be required to determine whether competition is involved in the activity-dependent loss of synaptic NMDA
receptor clusters in spontaneously active neuronal cultures. At the neuromuscular junction, the effect of local
activity blockade is a change in both the presynaptic and
postsynaptic compartments. In our neuronal cultures,
activity had no detectable effect on innervation but
acted selectively on one postsynaptic receptor type.
Thus, at this stage, it is simplest to envision a purely
postsynaptic mechanism following the initial change in
activity.
Physiological Consequences of Activity-Driven
Changes in NMDA Receptor Clustering
What could be the physiological role of activity driving
NMDA receptors away from synapses and inactivity inducing a buildup of synaptic NMDA receptors? At first
glance, the results reported here may appear counterintuitive, but it is important to point out the time course
of these studies, that the activity effect is not apparent
within minutes (such as with long-term potentiation) but
only after days. We consider this effect a novel kind of
“metaplasticity,” a higher order form of synaptic plasticity manifest as a change in the ability to undergo subsequent synaptic plasticity (Abraham and Bear, 1996). The
NMDA receptor is required for many forms of long-term
potentiation as well as activity-dependent synaptic remodeling during development (Bliss and Collingridge,
1993; Scheetz and Constantine-Paton, 1994; Tsien et
al., 1996). Our results would predict that a neuron’s history of synaptic activity may modulate its NMDA receptor distribution and thus its capacity to undergo further
activity-dependent synaptic plasticity. Under extreme
conditions, this enhanced sensitivity may result in pathological consequences of seizure and excitotoxicity.
Long-term exposure of neurons to NMDA receptor
antagonists has been reported to enhance NMDA responsiveness, NMDA-dependent calcium influx, and excitotoxicity upon removal of blockade. Hippocampal
neurons grown for 0.5–5 months under conditions of
activity blockade and then returned to normal medium
generated intense seizure-like activity that led to massive cell death (Furshpan and Potter, 1989). An increase
in synaptic NMDA receptors as reported here might enhance synchronized activity in the cultures and thus lead
to epileptiform activity. Similarly, hypothalamic neurons
treated chronically with APV or CNQX showed dramatic
rises in internal calcium levels and synchronized calcium
spiking after block removal, followed by cell death
(Obrietan and van den Pol, 1995). Pretreatment with the
NMDA receptor antagonist MK-801 24 hr before NMDA
injection in vivo resulted in enhanced brain injury, coupled to increases in the amount of NMDA receptor measured by receptor binding (McDonald et al., 1990). In
our cultures, we found that removal of activity blockade
resulted in reversion of the neurons to the control distribution of NMDA receptors without cell death. This difference may be due to the comparatively low cell density
at which our cultures were grown and/or the younger
age at the time of reversal.
From studies of visual system development, it has
been suggested that activity through synaptic NMDA
receptors as they build up during development may lead
to their down-regulation as a means of preventing excitotoxicity and/or regulating the ability of the system to
undergo long-term synaptic modification (Scheetz and
Constantine-Paton, 1994). For example, during development of the visual cortex, there is a critical period of
maximal plasticity of synaptic organization that corresponds with the period of maximal NMDA responsiveness (Carmignato and Vicini, 1992). Dark rearing or
blockade of electrical activity by tetrodotoxin prolongs
the critical period by blocking the decrease in NMDA
responsiveness seen with maturation (Fox et al., 1991;
Carmignato and Vicini, 1992) The results we report here
constitute a potential cellular mechanism for these phenomena, where increases in overall network activity with
maturation could induce a shift of NMDA receptors away
from synapses, thus lowering the probability of excitotoxicity, while limited activity in the immature state could
allow NMDA receptors to accumulate at synapses, thus
prolonging the period of high NMDA responsiveness.
Neuron
810
An important question remaining is whether the activitydependent redistribution of NMDA receptors is a developmentally restricted phenomenon or whether it can
occur in a mature neuron.
Conclusion
These results suggest a novel mechanism of modulation
of neuronal function by regulation of the subcellular distribution of the NMDA-type glutamate receptor ion channel by activity. Regulation of NMDA receptor clustering
and synaptic localization may permit neurons to change
their NMDA responsiveness and thus their ability to undergo activity-dependent synaptic change and/or excitotoxic damage.
Experimental Procedures
Cell Culture
Hippocampal neuronal cultures were prepared from 18 day embryonic rats as in Goslin and Banker (1991). In brief, hippocampi were
dissociated by trypsin treatment and trituration and plated on polyL-lysine-coated glass coverslips at a cell density of 2,400/cm 2. Neurons were maintained in MEM with N2 supplements above a glial
feeder layer. Similar results were found with neurons prepared from
postnatal rats and cultured in the presence of serum. Neurons were
treated for 1–40 days, with renewal every 3–4 days, with: 100 mM
APV, 1 mM CNQX, 5 mM nifedipine, 1 mM TTX, or 1 mM TTX plus 5
mM NMDA. The neurons were fixed 2–42 days after plating and
used for immunocytochemistry. Unless stated otherwise, all images
shown here and used for quantitation were from cells treated for 7
days (CNQX and nifedipine) or 7–14 days (APV and TTX) and fixed
at 16–22 days.
Immunocytochemistry and Quantitation
For NMDA receptor immunostaining, neurons were fixed with methanol for 10 min at 2208C, blocked with 10% BSA, and incubated
with mouse monoclonal antibody 54.1 to NMDAR1 (PharMingen;
0.1–3 mg/ml). There was a large amount of variation in staining
intensity between different lots of the antibody obtained from the
manufacturer, but the specificity of all batches was the same in
Western blots. Double- and triple-label immunostaining was done
with anti-NR1 and combinations of: rabbit anti-NR2B antiserum (Upstate Biotechnology, 4.2 mg/ml), rabbit anti-NR2A antiserum (gift of
M. Sheng, 1 mg/ml), guinea pig anti-PSD-95 antiserum (gift of M.
Sheng, 1 mg/ml), guinea pig anti-GluR1 antiserum (gift of R. L. Huganir, 1:1600), rabbit antiserum against synaptophysin (G95, gift
of P. DeCamilli; 1:8000), and appropriate fluorochrome-conjugated
secondary antibodies (Vector Labs, 2.5–7.5 mg/ml). Methanol fixation was required to obtain specific staining for NR1, NR2A, and
NR2B for reasons that are not clear; methanol may induce a conformational change or extract other masking proteins to uncover antigenic sites. The necessity for methanol fixation did not allow us to
distinguish surface from intracellular receptors. The sensitivity of
GluR1 staining was not decreased by methanol fixation (compare
Figures 4 and 6). The number of NR1 and GluR1 clusters was variable
between cultures but was consistent within a culture. The specificity
of all antibodies was confirmed on Western blots of culture extracts
(e.g., see Figure 7). The similar staining patterns with the independent antibodies against NR1, NR2A, and NR2B also serve as an
internal control for specificity of NMDA receptor staining. GluR1
immunostaining with paraformaldehyde fixation was as described
in Craig et al. (1994). Coverslips were mounted in Tris-HCl, glycerol,
polyvinyl alcohol with 2% 1,4-diazabicyclo[2,2,2]octane. Fluorescent images of cells were captured on a Photometrics series 200
cooled CCD camera mounted on a Zeiss Axioskop microscope with
a 633, 1.4 N.A. lens using Oncor imaging software.
For quantitation, CCD images were background subtracted, flat
field divided, and interactively thresholded to define clusters. A single threshold was chosen manually for each image so that NR1
clusters corresponded to spots of about 2-fold or greater intensity
above the diffuse fluorescence on the dendritic shaft. Each NR1
cluster was classified as synaptic or nonsynaptic based on evaluation of an overlay of the thresholded NR1 and synaptophysin images.
For the quantitative comparison of NR1 and GluR1 clusters, NR1
clusters were selected by thresholding, and GluR1 clusters were
selected by thresholding in combination with selecting regions of
interest to exclude thick dendrite shafts; the diffuse staining for
GluR1 was high within dendrite shafts, and so potential shaft clusters could not be reliably visualized. For ratio imaging of NMDA
receptor subunits, a mask was made for each cell corresponding
to NR1 clusters from the control cells representing the proximal and
distal largely nonsynaptic shaft clustering patterns or from the APVtreated cells representing the spiny synaptic pattern, and the mask
was then applied to both NR1 and NR2A/B images to obtain fluorescent intensity measurements. Measurements were analyzed using
Microsoft Excel, StatView, and CricketGraph. Images were prepared
for printing using Adobe Photoshop.
Western Blot Analysis
Cells from hippocampal cultures plated at high density (14,300/cm2 )
and between 15 and 17 days in culture were scraped into warm
phosphate-buffered saline, pelleted, and resuspended in Laemmli
buffer. Test coverslips from the same dishes used for the Western
blot analysis were fixed and immunostained for NR1 and synaptophysin to confirm the differential localization of NR1 in the control
compared to APV-treated groups. After SDS-PAGE and blotting
onto nitrocellulose, paired lanes of control and APV-treated samples
were probed with antibodies to NR1 (Pharmingen, 0.5 mg/ml), NR1N1, NR1-C1, NR2A, and NR2B (gifts of M. Sheng, all used at 1 mg/ml
as in Sheng et al. [1994]). HRP-conjugated secondary goat antirabbit or anti-mouse antisera (Jackson Labs) were used at dilutions
of 1:5000, and the signal was visualized using chemiluminescent
Super-Signal HRP substrate (Pierce Biochemicals) to expose XAR-5
X-ray film. The blots were stripped and reprobed with an anti-tubulin
antibody (1:20,000, DM1a, gift of V. Gelfand) and the signal visualized as before. The film signals were digitally scanned, and the
signal on the digital image was quantitated using NIH-Image.
Acknowledgments
We thank Anna S. Serpinskaya for excellent technical assistance,
M. Sheng, P. DeCamilli, R. L. Huganir, and V. Gelfand for gifts of
antibodies, and G. Banker, C. R. Bramham, C. Q. Doe, M. Gillette,
M. Sheng, G. S. Withers, and members of the Craig lab for comments
on a previous version of the manuscript. This work was supported
by the Markey Charitable Trust and NIH grant NS33184 to A. M. C.
Received June 3, 1997; revised August 27, 1997.
References
Abraham, W.C., and Bear, M.F. (1996). Metaplasticity: the plasticity
of synaptic plasticity. Trends Neurosci. 19, 126–130.
Aoki, C., Venkatesan, C., Go, C.-G., Mong, J.A., and Dawson, T.M.
(1994). Cellular and subcellular localization of NMDA-R1 subunit
immunoreactivity in the visual cortex of adult and neonatal rats. J.
Neurosci. 14, 5202–5222.
Bading, H., Ginty, D., and Greenberg, M.E. (1993). Regulation of gene
expression in hippocampal neurons by distinct calcium signaling
pathways. Science 260, 181–186.
Balice-Gordon, R.J., and Lichtman, J.W. (1994). Long term synapse
loss induced by focal blockade of postsynaptic receptors. Nature
372, 519–524.
Basarsky, T.A., Parpura, V., and Haydon, P.G. (1994). Hippocampal
synaptogenesis in cell culture: developmental time course of synapse formation, calcium influx, and synaptic protein distribution. J.
Neurosci. 14, 6402–6411.
Bekkers, J.M., and Stevens, C.F. (1989). NMDA and non-NMDA receptors are co-localized at individual excitatory synapses in cultured
rat hippocampus. Nature 341, 230–233.
Activity Effect on NMDA Receptor Clustering
811
Bessho, Y., Nawa, H., and Nakanishi, S. (1994). Selective up-regulation of an NMDA receptor subunit mRNA in cultured cerebellar granule cells by K1-induced depolarization and NMDA treatment. Neuron
12, 87–95.
Bliss, T.V.P., and Collingridge, G.L. (1993). A synaptic model of
memory: long-term potentiation in the hippocampus. Nature 361,
31–39.
Carmignato, G., and Vicini, S. (1992). Activity-dependent decrease in
NMDA receptor responses during development of the visual cortex.
Science 258, 1007–1011.
Cho, K.O., Hunt, C.A., and Kennedy, M.B. (1992). The rat brain postsynaptic density fraction contains a homolog of the Drosophila
discs-large tumor suppressor protein. Neuron 9, 929–942.
Choi, D.W. (1994). Glutamate receptors and the induction of excitotoxic neuronal death. Prog. Brain Res. 100, 47–51.
Cohen, N.A., Brenman, J.E., Snyder, S.H., and Bredt, D.S. (1996).
Binding of the inward rectifier K 1 channel Kir 2.3 to PSD-95 is regulated by protein kinase A phosphorylation. Neuron 17, 759–767.
Craig, A.M., Blackstone, C.D., Huganir, R.L., and Banker, G. (1993).
The distribution of glutamate receptors in cultured rat hippocampal
neurons: postsynaptic clustering of AMPA-selective subunits. Neuron 10, 1055–1068.
Craig, A.M., Blackstone, C.D., Huganir, R.L., and Banker, G. (1994).
Selective clustering of glutamate and g-aminobutyric acid receptors
opposite terminals releasing the corresponding neurotransmitters.
Proc. Natl. Acad. Sci. USA 91, 12373–12377.
Debski, E.A., Cline, H.T., McDonald, J.W., and Constantine-Paton,
M. (1991). Chronic application of NMDA decreases the NMDA sensitivity of the evoked tectal potential in the frog. J. Neurosci. 11,
2947–2957.
Didier, M., Mienville, J.M., Soubrie, P., Bockaert, J., Berman, S.,
Bursztajn, S., and Pin, J.P. (1994). Plasticity of NMDA receptor expression during mouse cerebellar granule cell development. Eur. J.
Neurosci. 6, 1536–1543.
Duclert, A., and Changeaux, J.-P. (1995). Acetylcholine receptor
gene expression at the developing neuromuscular junction. Physiol.
Rev. 75, 339–368.
Durand, G.M., Bennett, M.V., and Zukin, R.S. (1993). Splice variants
of the N-methyl-D-aspartate receptor NR1 identify domains involved
in regulation by polyamines and protein kinase. Proc. Natl. Acad.
Sci. USA 90, 6731–6735.
Ehlers, M.D., Tingley, W.G., and Huganir, R.L. (1995). Regulated
subcellular distribution of the NR1 subunit of the NMDA receptor.
Science 269, 1734–1737.
Feldmeyer, D., and Cull-Candy, S. (1996). Functional consequences
of changes in NMDA receptor subunit expression during development. J. Neurocytol. 25, 857–867.
Fletcher, T.L., Cameron, P., DeCamilli, P., and Banker, G. (1991).
Synaptogenesis in hippocampal cultures: evidence indicating that
axons and dendrites become competent to form synapses at different stages of neuronal development. J. Neurosci. 11, 1617–1628.
Follesa, P., and Ticku, M.N. (1996). NMDA receptor upregulation:
molecular studies in cultured mouse cortical neurons after chronic
antagonist exposure. J. Neurosci. 16, 2172–2178.
Forrest, D., Yuzaki, M., Soares, H.D., Ng, L., Luk, D.C., Sheng, M.,
Stewart, C.L., Morgan, J.I., Connor, G.A., and Curran, T. (1994).
Targeted disruption of NMDA receptor 1 gene abolishes NMDA
response and results in neonatal death. Neuron 13, 325–328.
Fox, K., Daw, N., Sato, H., and Czepita, D. (1991). Dark-rearing delays
the loss of NMDA-receptor function in kitten visual cortex. Nature
350, 342–344.
Furshpan, E.J., and Potter, D.D. (1989). Seizure-like activity and
cellular damage in rat hippocampal neurons in cell culture. Neuron
3, 199–207.
Ghosh, A., and Greenberg, M.E. (1995). Calcium signaling in neurons:
molecular mechanisms and cellular consequences. Science 268,
239–247.
Gomperts, S.N. (1996). Clustering membrane proteins: it’s all coming
together with the PSD-95/SAP90 protein family. Cell 84, 659–662.
Goslin, K., and Banker, G. (1991). Rat hippocampal neurons in low
density culture. In Culturing Nerve Cells, G. Banker and K. Goslin,
eds. (Cambridge, MA: MIT Press), pp. 251–282.
Hall, Z.W., and Sanes, J.R. (1993). Synaptic structure and development: the neuromuscular junction. Cell 72/Neuron 10 Suppl., 99–122.
Hall, R.A., and Soderling, T.R. (1997). Differential surface expression
and phosphorylation of the N-methyl-D-aspartate receptor subunits
NR1 and NR2 in cultured hippocampal neurons. J. Biol. Chem. 272,
4135–4140.
Huntley, G.W., Vickers, J.C., Brose, N., Janssen, W., Heinemann,
S.H., and Morrison, J.H. (1994). Distribution and synaptic localization
of immunocytochemically identified NMDA receptor subunit proteins in sensorimotor and visual cortices of monkey and human. J.
Neurosci. 14, 3603–3619.
Isaac, J.T.R., Nicoll, R.A., and Malenka, R.C. (1995). Evidence for
silent synapses: implications for the expression of LTP. Neuron 15,
427–434.
Ishii, T., Moriyoshi, K., Sugihara, H., Sakurada, K., Kadotani, H.,
Yokoi, M., Akazawa, C., Shigemoto, R., Mizuno, M., Masu, M., and
Nakanishi, S. (1993). Molecular characterization of the family of the
N-methyl-D-aspartate receptor subunits. J. Biol. Chem. 268, 2836–
2843.
Jones, K., and Baughman, R.W. (1991). Both NMDA and non-NMDA
subtypes of glutamate receptors are concentrated at synapses on
cerebral cortical neurons in culture. Neuron 17, 103–113.
Johnson, R.R., Jiang, X., and Burkhalter, A. (1996). Regional and
laminar differences in synaptic localization of NMDA receptor subunit NR1 splice variants in rat visula cortex and hippocampus. J.
Comp. Neurol. 368, 335–355.
Kharazia, V.N., Phend, K.D., Rustioni, A., and Weinberg, R.J. (1996).
EM colocalization of AMPA and NMDA receptor subunits at synapses in rat cerebral cortex. Neurosci. Lett. 210, 37–40.
Kim, E., Niethammer, M., Rothschild, A., Jan, Y.N., and Sheng, M.
(1995). Clustering of Shaker-type K1 channels by interaction with
a family of membrane associated guanylate kinases. Nature 378,
85–88.
Kim, E., Cho, K.O., Rothschild, A., and Sheng, M. (1996). Heteromultimerization and NMDA receptor-clustering activity of chapsyn-110,
a member of the PSD-95 family of proteins. Neuron 17, 103–113.
Kistner, U., Wenzel, B.M., Veh, R.W., Cases-Langhoff, C., Garner,
A.M., Appeltauer, U., Voss, B., Gundelfinger, E.D., and Garner, C.C.
(1993). SAP90, a rat presynaptic protein related to the product of
the Drosophila tumor supressor gene dlg-A. J. Biol. Chem. 268,
4580–4583.
Kornau, H.-C., Schenker, L.T., Kennedy, M.B., and Seeburg, P.H.
(1995). Domain interaction between NMDA receptor subunits and
the postsynaptic density protein PSD-95. Science 269, 1737–1740.
Kullmann, D.M., and Siegelbaum, S.A. (1995). The site of expression
of NMDA receptor-dependent LTP: new fuel for an old fire. Neuron
15, 997–1002.
Kutsuwada, T., Kashiwabuchi, N., Mori, H., Sakimura, K., Kushiya,
E., Araki, K., Meguro, H., Masaki, H., Kumanishi, T., Arakawa, M.,
and Mishina, M. (1992). Molecular diversity of the NMDA receptor
channel. Nature 358, 36–41.
Laurie, D.J., and Seeburg, P.H. (1994). Regional and developmental
heterogeneity in splicing of the rat brain NMDAR1 mRNA. J. Neurosci. 14, 3180–3194.
Li, Y., Erzurumlu, R., Chen, C., Zhaveri, S., and Tonegawa, S. (1994).
Whisker-related neuronal patterns fail to develop in the trigeminal
brainstem nuclei of NMDAR1 knockout mice. Cell 76, 427–437.
Liao, D., Hessler, N.A., and Malinow, R. (1995). Activation of postsynaptically silent synapses during pairing-induced LTP in CA1 region
of hippocampal slice. Nature 375, 400–403.
McBain, C.J., and Mayer, M.L. (1994). N-Methyl-D-aspartic acid receptor structure and function. Physiol. Rev. 74, 723–760.
McDonald, J.W., Silverstein, F.S., and Johnston, M.V. (1990). MK801 pretreatment enhances N-methyl-D-aspartate-mediated brain
injury and increases brain N-methyl-D-aspartate recognition site
binding in rats. Neuroscience 38, 103–113.
McIlhinney, R.A.J., Molnar, E., Atack, J.R., and Whiting, P.J. (1996).
Neuron
812
Cell surface expression of the human N-methyl-D-aspartate receptor subunit 1a requires the co-expression of the NR2a subunit in
transfected cells. Neuroscience 70, 989–997.
Monyer, H., Sprengel, R., Schoepfer, R., Herb, A., Higuchi, M., Lomeli, H., Burnashev, N., Sakmann, B., and Seeburg, P.H. (1992).
Heteromeric NMDA receptors: molecular and functional distinction
of subtypes. Science 256, 1217–1221.
Monyer, H., Burnashev, N., Laurie, D.J., Sakmann, B., and Seeburg,
P.H. (1994). Developmental and regional expression in the rat brain
and functional properties of four NMDA receptors. Neuron 12,
529–540.
Moriyoshi, K., Masu, M., Ishii, T., Shigemoto, R., Mizuno, N., and
Nakanishi, S. (1991). Molecular cloning and characterization of the
rat NMDA receptor. Nature 354, 31–37.
Nguyen, Q.T., and Lichtman, J.W. (1996). Mechanism of synapse
disassembly at the developing neuromuscular junction. Curr. Opin.
Neurobiol. 6, 104–112.
Niethammer, M., Kim, E., and Sheng, M. (1996). Interaction between
the C terminus of NMDA receptor subunits and multiple members
of the PSD-95 family of membrane-associated guanylate kinases.
J. Neurosci. 16, 2157–2163.
Obrietan, K., and van den Pol, A.N. (1995). Calcium hyperexcitability
in neurons cultured with glutamate receptor blockade. J. Neurophysiol. 73, 1524–1536.
Petralia, R.S., Yokotani, N., and Wenthold, R.J. (1994a). Light and
electron microscope distribution of the NMDA receptor subunit
NMDAR1 in the rat nervous system using a selective anti-peptide
antibody. J. Neurosci. 14, 667–696.
Petralia, R.S., Wang, Y.-X., and Wenthold, R.J. (1994b). The NMDA
receptor subunits NR2A and NR2B show histological and ultrastructural localization patterns similar to those of NR1. J. Neurosci. 14,
6102–6120.
Resink, A., Villa, M., Benke, D., Mohler, H., and Balazs, R. (1995).
Regulation of the expression of NMDA receptor subunits in rat cerebellar granule cells: effect of chronic K1-induced depolarization and
NMDA exposure. J. Neurochem. 64, 558–565.
Scheetz, A.J., and Constantine-Paton, M. (1994). Modulation of
NMDA receptor function: implications for vertebrate neural development. FASEB J. 8, 744–752.
Segal, M. (1995). Imaging of calcium variations in living dendritic
spines of cultured rat hippocampal neurons. J. Physiol. 486,
283–295.
Sheng, M., Cummings, J., Roldan, L.A., Jan, Y.N., and Jan, L.Y.
(1994). Changing composition of heteromeric NMDA receptors during development of rat cortex. Nature 368, 144–147.
Siegel, S.J., Brose, N., Janssen, W.G., Gasic, G.P., Jahn, R., and
Heinemann, S.F. (1994). Regional, cellular, and ultrastructural distribution of N-methyl-D-aspartate receptor subunit 1 in monkey hippocampus. Proc. Natl. Acad. Sci. USA 91, 564–568.
Siegel, S.J., Janssen, W.G., Tullai, J.W., Rogers, S.W., Moran, T.,
Heinemann, S.F., and Morrison, J.H. (1995). Distribution of the excitatory amino acid receptor subunits GluR2(4) in monkey hippocampus and colocalization with subunits GluR5–7 and NMDAR1. J. Neurosci. 15, 2707–2719.
Tsien, J.Z., Huerta, P.T., and Tonegawa, S. (1996). The essential role
of hippocampal CA1 NMDA receptor-dependant synaptic plasticity
in spatial memory. Cell 87, 1317–1326.
Vallano, M.L., Lambolez, B., Audinat, E., and Rossier, J. (1996).
Neuronal activity differentially regulates NMDA receptor subunit expression in cerebellar granule cells. J. Neurosci. 16, 631–639.
Williams, K., Dichter, M.A., and Molinoff, P.B. (1992). Up-regulation
of N-methyl-D-aspartate receptors on cultured cortical neurons after
exposure to antagonists. Mol. Pharmacol. 42, 147–151.
Wyszynski, M., Lin, J., Rao, A., Nigh, E., Beggs, A.H., Craig, A.M., and
Sheng, M. (1997). Competitive binding of a-actinin and calmodulin to
the NMDA receptor. Nature 385, 439–442.