ArticlePDF Available

A Ca2+-Based Computational Model for NMDA Receptor-Dependent Synaptic Plasticity at Individual Post-Synaptic Spines in the Hippocampus

Authors:

Abstract and Figures

Associative synaptic plasticity is synapse specific and requires coincident activity in pre-synaptic and post-synaptic neurons to activate NMDA receptors (NMDARs). The resultant Ca(2+) influx is the critical trigger for the induction of synaptic plasticity. Given its centrality for the induction of synaptic plasticity, a model for NMDAR activation incorporating the timing of pre-synaptic glutamate release and post-synaptic depolarization by back-propagating action potentials could potentially predict the pre- and post-synaptic spike patterns required to induce synaptic plasticity. We have developed such a model by incorporating currently available data on the timecourse and amplitude of the post-synaptic membrane potential within individual spines. We couple this with data on the kinetics of synaptic NMDARs and then use the model to predict the continuous spine [Ca(2+)] in response to regular or irregular pre- and post-synaptic spike patterns. We then incorporate experimental data from synaptic plasticity induction protocols by regular activity patterns to couple the predicted local peak [Ca(2+)] to changes in synaptic strength. We find that our model accurately describes [Ca(2+)] in dendritic spines resulting from NMDAR activation during pre-synaptic and post-synaptic activity when compared to previous experimental observations. The model also replicates the experimentally determined plasticity outcome of regular and irregular spike patterns when applied to a single synapse. This model could therefore be used to predict the induction of synaptic plasticity under a variety of experimental conditions and spike patterns.
This content is subject to copyright.
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 1
SYNAPTIC NEUROSCIENC
E
Original research article
published: 21 July 2010
doi: 10.3389/fnsyn.2010.00031
to vary Ca2+ influx through NMDARs by fixing the post-synaptic
membrane potential during low frequency synaptic stimulation
(Isaac et al., 1995; Daw et al., 2000).
Spike timing-dependent plasticity (STDP) is a form of Hebbian
synaptic plasticity that incorporates a temporal specificity to coin-
cident pre- and post-synaptic activity. In the hippocampus, STDP
was originally thought to be induced by single pairs of pre- and
post-synaptic action potentials such that if the pre-synaptic action
potential occurs before the post-synaptic action potential LTP is
induced whereas if the order of action potentials is reversed then
LTD is induced (Bi and Poo, 1998; Debanne et al., 1998; Nishiyama
et al., 2000; Campanac and Debanne, 2008; Kwag and Paulsen,
2009). Other data have proposed this model should include a
requirement for bursts of post-synaptic action potentials for the
induction of LTP although LTD may be induced by single pairs
(Pike et al., 1999; Wittenberg and Wang, 2006; Buchanan and
Mellor, 2007) reviewed in (Buchanan and Mellor, 2010). This is
a divergence from the situation at cortical synapses where single
pairs of action potentials can induce both LTP and LTD (Sjostrom
et al., 2001; Sjostrom and Nelson, 2002; Froemke et al., 2006, but
see Nevian and Sakmann, 2006).
IntroductIon
Hebbian synaptic plasticity is the cellular and molecular correlate
of associative learning in the brain. During presentation of infor-
mation that needs to be retained for future use, specific synapses
are subjected to activity patterns that induce a long-term change
in synaptic strength. For Hebbian synaptic plasticity at Schaffer
collateral synapses in the hippocampus, these patterns require
coincident activity in pre- and post-synaptic neurons to activate
NMDA receptors (NMDARs) present on the membrane of the
post-synaptic dendritic spine. The resulting Ca2+ influx through
NMDARs is the critical trigger for induction of synapse specific
plasticity (Lisman, 1989).
Classically, high frequency synaptic stimulation induces long-
term potentiation (LTP) whereas low frequency stimulation induces
long-term depression (LTD) suggesting that brief high concentra-
tions of Ca2+ in the post-synaptic spine induce LTP whereas pro-
longed lower concentrations of Ca2+ induce LTD (Bear et al., 1987;
Hansel et al., 1996). This hypothesis is supported by measurements
of Ca2+ concentration during plasticity induction (Hansel et al.,
1997; Cho et al., 2001; Cormier et al., 2001; Ismailov et al., 2004;
Gall et al., 2005) and by plasticity induction protocols designed
A Ca2+-based computational model for NMDA receptor-
dependent synaptic plasticity at individual post-synaptic
spines in the hippocampus
Owen J. L. Rackham1, Krasimira Tsaneva-Atanasova 2, Ayalvadi Ganesh3 and Jack R. Mellor4*
1 Department of Engineering Mathematics, Bristol Centre for Complexity Sciences, University of Bristol, University Walk, Bristol, UK
2 Department of Engineering Mathematics, Bristol Centre for Applied Nonlinear Mathematics, University of Bristol, University Walk, Bristol, UK
3 Department of Mathematics, Bristol Centre for Complexity Sciences, University of Bristol, University Walk, Bristol, UK
4 Department of Anatomy, Medical Research Council Centre for Synaptic Plasticity, University of Bristol, University Walk, Bristol, UK
Associative synaptic plasticity is synapse specific and requires coincident activity in pre-synaptic
and post-synaptic neurons to activate NMDA receptors (NMDARs). The resultant Ca2+ influx is
the critical trigger for the induction of synaptic plasticity. Given its centrality for the induction
of synaptic plasticity, a model for NMDAR activation incorporating the timing of pre-synaptic
glutamate release and post-synaptic depolarization by back-propagating action potentials could
potentially predict the pre- and post-synaptic spike patterns required to induce synaptic plasticity.
We have developed such a model by incorporating currently available data on the timecourse
and amplitude of the post-synaptic membrane potential within individual spines. We couple this
with data on the kinetics of synaptic NMDARs and then use the model to predict the continuous
spine [Ca2+] in response to regular or irregular pre- and post-synaptic spike patterns. We then
incorporate experimental data from synaptic plasticity induction protocols by regular activity
patterns to couple the predicted local peak [Ca2+] to changes in synaptic strength. We find that
our model accurately describes [Ca2+] in dendritic spines resulting from NMDAR activation during
pre-synaptic and post-synaptic activity when compared to previous experimental observations.
The model also replicates the experimentally determined plasticity outcome of regular and
irregular spike patterns when applied to a single synapse. This model could therefore be used
to predict the induction of synaptic plasticity under a variety of experimental conditions and
spike patterns.
Keywords: synaptic plasticity, hippocampus, dendritic spines, NMDA receptor, spike timing-dependent plasticity
Edited by:
Per Jesper Sjöström, University
College London, UK
Reviewed by:
Harel Z. Shouval, University of Texas
Medical School at Houston, USA
Thomas G. Oertner, Friedrich Miescher
Institute for Biomedical Research,
Switzerland
*Correspondence:
Jack R. Mellor, Department of
Anatomy, Medical Research Council
Centre for Synaptic Plasticity,
University of Bristol, University Walk,
Bristol, UK. e-mail: jack.mellor@bristol.
ac.uk
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 2
Rackham et al. Model for hippocampal synaptic plasticity
Since Ca2+ influx through NMDARs is pivotal for LTP and LTD,
this suggests the induction of synaptic plasticity can be predicted
by NMDAR opening kinetics in response to pre-synaptic glutamate
release and post-synaptic depolarization. This approach has been
adopted for the modeling of post-synaptic calcium dynamics in
response to synaptic stimulation or back-propagating action poten-
tials (Franks et al., 2002; Grunditz et al., 2008; Keller et al., 2008)
and to STDP induction protocols (Shouval et al., 2002; Rubin et al.,
2005; Graupner and Brunel, 2007; Helias et al., 2008; Urakubo et al.,
2008; Castellani et al., 2009). However, these STDP models are limited
by the experimental data used to determine their parameters and,
in addition, ought to accurately predict the plasticity outcomes of
a variety of induction protocols. Recent advances in dendritic spine
imaging provide data on spine depolarization and Ca2+ concentrations
in response to pre- and post-synaptic action potentials (Sabatini et al.,
2002; Nevian and Sakmann, 2006; Bloodgood and Sabatini, 2007;
Canepari et al., 2007; Palmer and Stuart, 2009) that potentially greatly
increase the accuracy of such models of plasticity induction.
We have developed a computational model of synaptic plasticity
induction based on one originally described by Shouval et al. Our
model incorporates the latest experimental data on dendritic spine
depolarization and Ca2+ dynamics. We also test the predictive power
of our model on many plasticity induction protocols by calculating
continuous Ca2+ concentrations during long induction periods. We
find that our model accurately predicts the experimental data tested
and we hypothesize that it can thus be used to search for instances of
synaptic plasticity induction during continuous activity at Hebbian
synapses in the hippocampus.
MaterIals and Methods
We use a physiologically plausible model based on intracellular Ca2+
dynamics caused by NMDAR activation during the induction of syn-
aptic plasticity to predict the plasticity outcome of any set of pre- and
post-synaptic activity patterns that occur at the Schaffer collateral
synapse in the hippocampus. Since we are interested in studying
experimental spike trains we modify a model originally proposed by
Shouval et al. (2002) to allow us to carry out such analysis. We make
a number of critical modifications to analyze the Ca2+ dynamics in
individual dendritic spines during long periods of irregular spiking
activity. This is illustrated using short epochs of overlapping hip-
pocampal place cell activity (Figure 1) (Isaac et al., 2009).
Essential components of the experimental spike trains for the
activation of NMDARs and therefore the induction of synaptic
plasticity are (i) the pre-synaptic release of glutamate that dic-
tates the binding of glutamate to the NMDARs and (ii) the post-
synaptic membrane potential that determines the relative blockade
of NMDARs by Mg2+. For the purposes of this model, we have
assumed the two events that determine the post-synaptic mem-
brane potential within an individual dendritic spine are excita-
tory post-synaptic potentials (EPSPs) and back-propagating action
potentials (BPAPs). We start by modeling the BPAPs as follows:
BPAP ,
max
bs
f
bs i
post
f
bs
i
post
() exp()
tVItt
I
tt
=−−
+
<
τ
ss
bs i
post
s
bs
exp()−−
tt
τ
(1)
FIGURE 1 | Calculating [Ca2+] in dendritic spines during continuous pre-
and post-synaptic activity. The model initially calculates the membrane
potential during continuous activity by summating the membrane potential
changes due to EPSPs and BPAPs from a resting membrane potential of
65 mV. The Ca2+ current passing through synaptic NMDARs is then calculated
from the membrane potential and glutamate binding kinetics. Finally spine
[Ca2+] is calculated from Ca2+ buffering and diffusion kinetics. Left hand panels
show the post-synaptic responses to an epoch of place cell activity spanning
3500 ms. Right hand panels show a 200 ms excerpt from this epoch.
where
Vmax
bs
is the maximum depolarization due to the BPAP,
If
bs
, and
Is
bs
are the relative magnitudes of the fast and slow components of
the BPAP, respectively, that sum to one, and the integration time step
δ is 0.1 ms. Due to the slower (and much smaller) after- depolarizing
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 3
Rackham et al. Model for hippocampal synaptic plasticity
EPSPNMDA f
f
ep s
s
ep
m
() expexp (
tI tItVV
=
+
ττ
rr1
rest
NMDA
f
f
ep
f
ep s
s
ep
EPSP
)
() expex
V
dt
dt ItI
=−
11
τττpp ()
max( ). (max
=⇒
=⇒
tVV
V
t
τs
ep
mr1
rest
EPSP
0
00924 (()).
..
t
Nn
=⇒
=
00812
5
00812
If
and
are the relative magnitudes of the fast and slow component
of the NMDAR current as a result of glutamate binding, respec-
tively, that sum to one, and Θ is the Heaviside (unit) step function.
The voltage dependence of the current that takes into account Mg2+
block of the receptor (Jahr and Stevens, 1990) is represented by
the term B(Vm)(VmVr1)/Vrest, where (VmVr1) is the driving force
determined by the reversal potential, Vr1 (0 mV), and
BV KV
() exp( )([ ]/ .)
.
m
Mm Mg
=+−
1
1357
We then calculate the spine membrane potential as the summa-
tion of BPAP and EPSPAMPA and EPSPNMDA:
Vt Vt tt
mrestAMPANMDA
BPAP EPSP EPSP() () () (),=+ ++
(4)
where Vrest is set at 65 mV unless otherwise stated. An example
of the predicted spine voltage can be seen for a sample epoch of
overlapping place cell activity in Figure 1.
Since NMDARs provide the major source of Ca2+ influx into
post-synaptic dendritic spines (Bloodgood and Sabatini, 2007), we
incorporate in our model the Ca2+ current through NMDAR that
takes the following form (Shouval et al., 2002):
It PG It tt
I
i
tt
NMDA NMDA f
i
f
s
i
pre
() ()exp()
(
=−−
+
<
0Θ
Θ
τ
tt tt BV VV
i)exp ()
()(),
−−
i
s
mm r2
τ
(5)
This is similar to Eq. 3 except for the terms P0 and GNMDA that
represent the open channel probability and NMDAR Ca2+ con-
ductance respectively and Vr2 is the reversal potential for calcium
(130 mV).
Next, the rate of change of the [Ca2+] inside the post-synaptic
spine is governed by:
d
dt I
[] []
,
Ca Ca
NMDA
Ca
22++
=−ατ
(6)
where α is a factor that converts current to flux and τCa is the calcium
passive decay time constant. An example of the Ca2+ current flow
through NMDARs and the resulting predicted [Ca2+] in the spine
can be seen in Figure 1.
Finally, we assume that spine [Ca2+] is the trigger for synaptic
strength change. For the purposes of our study the continuous
model for synaptic strength used in Shouval et al. is modified to act
as a Ca2+-gated function based on local peaks in [Ca2+] as follows:
potential, if two spikes happen near enough to each other that the
first spike is still decaying, the effect of the BPAPs is additive. Since
we are modeling the BPAP at the spine
Vmax
bs
is set at 67 mV in line
with experimental data measuring membrane potential in spines
with voltage-sensitive dyes (Canepari et al., 2007; Palmer and Stuart,
2009). This is smaller than the maximum BPAP amplitude found at
the soma used by Shouval et al. An example of the modeled BPAP
during place cell activity can be seen in Figure 1.
The equation that governs the behavior of AMPAR-mediated
EPSPs in the model is similar to (1) having a slow and a fast expo-
nential component:
EPSPAMPA
a
i
pre
s
ep
i
pre
f
ep
()
exp()
exp()
t
Ntt tt
=−−
+−−
ττ
<
tt
VV
V
i
pre
mr1
rest
()
,
(2)
where the parameter Na reflects the maximum effect that a sin-
gle AMPAR-mediated EPSP can have. The value of Na can vary
depending on the number of synapses activated. Activation of a
single synapse results in a membrane depolarization in the spine of
approximately 10 mV (Palmer and Stuart, 2009). Again this deviates
from the value of 1 mV recorded at the soma and used by Shouval
et al. Assuming that the maximum depolarization that a single EPSP
can generate is 10 mV we define Na in the following way:
EPSPAMPA
f
ep
s
ep
mr1
() expexp ()
tttVV
V
=
+
ττ
rrest
AMPA
f
ep
f
ep
s
ep
s
ep
EPSP
dt
=−
dt tt() expexp
11
ττττ
=⇒
=− +
()
(log log)
VV
V
t
mr1
rest
f
ep
s
ep
s
ep
f
ep
f
0
ττ ττ
τeep
s
ep
EPSP
+
=⇒ =⇒
=
τ
max( ). (max()).
.
tt
Na
00128 06968
10
069668.
EPSPAMPA also depends on the membrane potential, Vm. This
dependence is represented by the term (VmVr1)/Vrest where Vr1 is
the reversal potential for AMPARs (0 mV) and Vrest is the resting
membrane potential (65 mV).
The equation that governs the behavior of NMDA-mediated
EPSPs in the model has the following form:
EPSPNMDA n
f
i
f
s
i
pre
() ()exp()
(
tN It tt
It
tt
i
i
=−−
+
<
Θ
Θ
τ
))exp ()
()
()
,
−−
tt BV VV
V
i
s
m
mr1
rest
τ
(3)
where the parameter Nn reflects the maximum effect of the
NMDAR-mediated component of the EPSP. This is calculated
in a similar fashion to Na for EPSPAMPA using a value of 5 mV
for the NMDAR-mediated EPSP at 65 mV in the absence of
Mg2+ measured by dendritic recordings (Fernandez de Sevilla
et al., 2007).
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 4
Rackham et al. Model for hippocampal synaptic plasticity
Table 1 | Parameter values of the synaptic model.
Parameter Value Parameter Value
If
bs
0.75 α1 0.3
τf
bs
3 ms α2 0.45
τs
bs
25 ms β1 80
τf
ep
5 ms β2 80
τs
ep
50 ms P1 100 ms
If 0.5 P2 0.02 ms
τf 50 ms P3 4
τs 200 ms P4 1000 ms
τCa 50 ms P0 0.5
Vmax
bs
67 mV GNMDA 0.002 μM/ms mV
Vrest 65 mV KM 0.092 mV1
Vr1 0 mV Na 14.35 mV
Vr2 130 mV Nn 61.58 mV
Having validated our model for the observed Ca2+ influx at
dendritic spines we next asked the question if the model could
replicate experimental data for the induction of synaptic plastic-
ity using a variety of protocols. We have restricted our model to
comparisons with experimental data from the Schaffer collateral
synapse of the hippocampus and not considered other synapses
in other brain regions.
spIke tIMIng-dependent plastIcIty wIth paIrs of pre- and
post-synaptIc spIkes
To model STDP with pairs of pre- and post-synaptic spikes we
initially assumed single synaptic activation and varied t between
20 and + 100 ms at intervals of 0.1 ms measuring the peak [Ca2+]
at each value of t (Figure 3A). [Ca2+] rose from its baseline of
72 nM (the peak [Ca2+] attained for a single EPSP in isolation) to
a peak of 230 nM at t 10 ms (Figure 3B). Experiments such as
these have been shown to generate no significant synaptic plasticity
(Buchanan and Mellor, 2007) whereas those using larger amplitude
EPSPs have been shown to generate LTD (Wittenberg and Wang,
2006). We estimated the activation of multiple synapses at the same
time would increase the depolarization within a single dendritic
spine during an EPSP from 10 to 20 mV based on experimental
predictions (Palmer and Stuart, 2009). Thus we have estimated that
the activation of other spines will contribute an additional 10 mV
of depolarization within an activated spine above and beyond the
experimentally determined 10 mV for activation of a single synapse.
This doubling of the EPSP amplitude resulted in an increase in
peak [Ca2+] at all values of t with a peak of 279 nM occurring at
t 10 ms (Figure 3C).
It has also been shown that the frequency of spike pairing is
important for the induction of plasticity such that at higher fre-
quencies (>5–10 Hz) LTP can be induced (Wittenberg and Wang,
2006; Buchanan and Mellor, 2007). We varied the frequency of
spike pairings in our model for 10 mV EPSPs over a range of
frequencies from 1 to 100 Hz (Figure 3D). Summation of Ca2+
transients was found to occur at frequencies greater than 5 Hz
indicating that increasing the frequency will shift the STDP pro-
tocol towards larger [Ca2+] and therefore LTP in line with the
experimental data.
Ca Ca :
Ca Ca
=∧ <
=
+
()()
++
+
++
[] []
[][],
22
1
22
00
1
W
WW
j
j
j
jj
ηΩ Ca 0
Ca Ca Ca
ΩΩ
[]
[][],[
j
jjjj
W
2
22 2
1
+
++ +
()
>
+
()()
η]]
()
0
(7)
where
[]
exp( ())
exp( ()).exp( (
Ca Ca
Ca
C
j
j
j
22
2
2
2
2
2
1
1025
+
+
+
=
+−
βα
βα
βaa
Ca
Ca Ca C
j
j
j
j
ff
2
1
1
2
1
2
2
1
1
+
+
+
+
+−
=
α
βα
η
))
exp( ()),
[]
()
,(
aa Ca
j
j
P
P
PP
21
2
24
3
+
+
=++)() .
The critical target for Ca2+ influx through NMDARs is the
enzyme CAMKII. Due to its ability to autophosphorylate, the acti-
vation of this molecule can be long lived and the level is determined
by local peak [Ca2+]. Thus, synaptic weight change is determined at
local peak [Ca2+] (Miller et al., 2005; Graupner and Brunel, 2007;
Helias et al., 2008; Urakubo et al., 2008; Castellani et al., 2009).
Since there is no noise associated with our model, these peaks are
measured instantaneously without smoothing. Experimentally,
increases in synaptic weight tend towards saturation as synaptic
weight increases. In addition, decreases require synaptic weight to
always be >0. These constraints explain the form of Eq. 7.
Numerical integration was performed using forward Euler
method implemented in MATLAB.
The parameter values used in the simulations are given for com-
pleteness in Table 1.
results
Our starting point for developing a model for the induction of
synaptic plasticity was to incorporate the most recent and accurate
measurements of voltage changes within dendritic spines using data
from measurements of voltage-dependent dyes (Canepari et al.,
2007; Palmer and Stuart, 2009). We model the membrane poten-
tial at the spine rather than the soma because this is the site of the
NMDARs critical for the induction of synaptic plasticity. This shifts
the determination of membrane depolarization away from BPAPs
and towards EPSPs since the former attenuate as they pass along the
dendrite and the latter are now measured at their site of origin. This
is a departure from previous models that used values for BPAPs and
EPSPs recorded at the soma (Shouval et al., 2002). With this change,
our model predicts that an EPSP resulting from the activation of a
single synapse is sufficient to cause a significant Ca2+ influx through
NMDARs (Figure 2A) in line with experimentally observed data
(Bloodgood and Sabatini, 2007; Canepari et al., 2007; Sobczyk and
Svoboda, 2007). The pairing of a BPAP with a single EPSP with a
time delay of 10 ms produces 3–4 times the Ca2+ influx (Figure 2A)
that again agrees qualitatively with experimentally observed data
(Bloodgood and Sabatini, 2007). For comparison we changed the
maximal EPSP and BPAP amplitudes to those known to occur at
the soma (1 and 100 mV respectively). With these parameters, a
single EPSP produces limited Ca2+ influx whereas pairing an EPSP
with a BPAP produces a large Ca2+ influx (Figure 2B).
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 5
Rackham et al. Model for hippocampal synaptic plasticity
plasticity is induced (Buchanan and Mellor, 2007). When we used
theta burst stimulation with five stimuli to only the pre-synaptic
input, the model predicted peak [Ca2+] within the spine to be
325 nM (Figure 5) and with four stimuli 250 nM. The value for
four stimuli is more physiologically relevant since the probability
of neurotransmitter release at any one Schaffer collateral synapse
is considerably less than 1. Therefore it is highly unlikely that
an experimental theta burst will ever generate five EPSPs at an
individual synapse.
paIrIng post-synaptIc depolarIzatIon wIth
pre-synaptIc stIMulatIon
Other common synaptic plasticity induction protocols have dis-
pensed with the need for post-synaptic spikes altogether and use
voltage clamp to depolarize the post-synaptic membrane and allow
NMDAR activation. This technique neatly demonstrates the bidi-
rectional nature of NMDAR-dependent plasticity since depolariza-
tion to moderate levels (40 mV) produces LTD whereas higher
depolarization (0 mV) produces LTP (Isaac et al., 1995; Daw et al.,
2000). We tested this with our model by clamping the membrane
potential (Vm) at either 40 or 0 mV (Figure 6). Peak [Ca2+] in
response to EPSPs were 336 nM and 2.43 μM respectively, which
when compared to peak [Ca2+] produced by other protocols would
be expected to induce LTD and LTP respectively in agreement with
experimental data.
the ca2+ hypothesIs can explaIn prevIous experIMental data
The Ca2+ hypothesis states that brief high concentrations of Ca2+
in the post-synaptic spine induce LTP whereas prolonged lower
concentrations of Ca2+ induce LTD (Bear et al., 1987; Hansel et al.,
1996). This is expressed graphically in Figure 7. Points are indi-
cated representing the predicted [Ca2+] from our model for specific
plasticity inducing protocols. STDP with single pairs of BPAPs and
small EPSPs do not induce plasticity (Buchanan and Mellor, 2007)
but when large EPSPs are used LTD is induced (Wittenberg and
Wang, 2006) and STDP with triplets of single EPSPs and bursts of
BPAPs produces LTP (Pike et al., 1999; Wittenberg and Wang, 2006;
Buchanan and Mellor, 2007). When the post-synaptic membrane
potential is set at 40 mV during pre-synaptic stimulation LTD is
induced (Daw et al., 2000) whereas at 0 mV LTP is induced (Isaac
et al., 1996). Theta burst pairing also induces LTP (Frick et al.,
spIke tIMIng-dependent plastIcIty wIth trIplets of spIkes
Post-synaptic burst firing has been shown to be important for the
induction of LTP at Schaffer collateral synapses in the hippocampus
where burst firing in this instance refers to any number of spikes
greater than one (Pike et al., 1999; Wittenberg and Wang, 2006;
Buchanan and Mellor, 2007). We tested this on our model using
triplets of spikes composed of one pre-synaptic spike and two post-
synaptic spikes where t is the time between the pre-synaptic spike
and the first post-synaptic spike and s is the delay between the two
post-synaptic spikes (Figure 4A). We first used 10 mV EPSPs with a
constant s of 10 ms and varied t between 20 and +100 ms. This
produced a peak [Ca2+] of 420 nM at t = 4 ms which increased to
a peak [Ca2+] of 475 nM when 20 mV EPSPs were used (Figure 4B)
confirming that spike triplets produce higher peak [Ca2+] than spike
pairs and therefore are more likely to induce LTP.
We next varied s whilst maintaining t constant at 10 ms for
both 10 and 20 mV EPSPs revealing a decrease in peak [Ca2+] as s
increases (Figure 4C). Finally, we varied the frequency of triplets
for 10 mV EPSPs over a range of frequencies from 1 to 100 Hz
whilst keeping t and s constant at 10 ms each (Figure 4D).
Summation of Ca2+ transients was found to occur at frequencies
greater than 4 Hz.
theta burst plastIcIty
We now moved away from STDP to look at other common syn-
aptic plasticity induction protocols. The theta burst protocol was
developed to mimic the activity patterns believed to occur at hip-
pocampal synapses during learning and consists of bursts of four
or five spikes at 100 Hz with an interburst interval of 200 ms.
These can either be applied to the pre- and post-synaptic neuron
coincidentally (Frick et al., 2004) or to just the pre-synaptic neuron
(Larson et al., 1986). The latter then leads to post-synaptic spikes
through EPSP summation if the initial EPSP amplitude is suf-
ficiently large (Buchanan and Mellor, 2007). We used our model
to mimic coincident theta burst activity in pre- and post-synaptic
neurons using 10 mV EPSPs and found that this type of synap-
tic stimulation produces very large peak [Ca2+] within dendritic
spines (Figure 5) indicating that this protocol is very efficient at
producing LTP in agreement with experimental data. Experimental
data also shows when theta burst stimulation is given to only the
pre-synaptic neuron without initiating action potentials then no
FIGURE 2 | Comparison of predicted [Ca2+] dynamics in dendritic spines and in the soma. [Ca2+] profiles in response to a 10 mV EPSP at the spine (A) or a 1 mV
EPSP at the soma (B) on their own (gray) or in combination with a BPAP (black) of amplitude 60 mV at the spine (A) or 100 mV at the soma (B). The delay between
EPSP and BPAP initiation is 10 ms.
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 6
Rackham et al. Model for hippocampal synaptic plasticity
InductIon of synaptIc plastIcIty by hIppocaMpal place cell
fIrIng patterns
One of the main purposes for developing a model that is capable
of continuously monitoring [Ca2+] in spines and therefore predicts
changes in synaptic strength is to scan long periods of neuronal activ-
ity for epochs that would be expected to induce plasticity without
having to directly measure synaptic strength. To test if the model could
perform this task we used data from experiments where long sections
of hippocampal place cell activity were replayed into single hippoc-
ampal synapses to test the plasticity outcome (Isaac et al., 2009).
2004; Buchanan and Mellor, 2007) whereas theta burst to only
pre-synaptic inputs does not (Buchanan and Mellor, 2007). In this
instance the absolute [Ca2+] values predicted by the model are not
as important as the relative magnitudes between plasticity induc-
tion protocols. However, it is interesting to note that the absolute
[Ca2+] values predicted by the model broadly agree with those
measured experimentally for the induction of synaptic plasticity at
Schaffer collateral synapses on CA1 pyramidal neurons (Cormier
et al., 2001). Thus the predictions from our model support the
Ca2+ hypothesis for synaptic plasticity induction.
FIGURE 3 | [Ca2+] dynamics in response to paired pre- and post-synaptic
spikes. (A) The model calculates [Ca2+] within a spine from the membrane
potential resulting from a pair of pre- and post-synaptic spikes. Gray line shows
EPSP in the absence of BPAP. Varying t shows that [Ca2+]max is greatest when
0 t 30 ms for 10 mV (B) or 20 mV (C) EPSPs. (D) The frequency of spike
pairings given at t = 10 ms determines [Ca2+]max.
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 7
Rackham et al. Model for hippocampal synaptic plasticity
robust LTP induction in agreement with the experimental data
(Figures 9A–D). We also tested two further pairs of place cells with
non-overlapping or adjacent place fields (1A and 1C, 2E and 2D) and
found the model predicted only a small LTD for the non- overlapping
pair and a small LTP for the adjacent pair (Figures 9E–F).
Our original experimental data also tested a pair of place cells
that had an asymmetric cross-correlation such that cell 1A prefer-
entially fired just before cell 1B. Because classical STDP rules state
that the temporal order of pre- and post-synaptic spikes controls
the direction of synaptic plasticity (Bi and Poo, 1998; Song et al.,
2000), the existence of this asymmetry suggested that when cell
1A was pre-synaptic and cell 1B post-synaptic then LTP would be
induced but if the cells were reversed then LTD would be induced.
We first took an 16-min period of activity from a pair of place
cells (1A and 1B) that had overlapping place fields and therefore
would be expected to fire at approximately the same time (Isaac
et al., 2009). During the 16-min period short coincident bursts of
activity could be seen in the two place cells that the model predicted
would produce large [Ca2+] sufficient to induce LTP (Figures 1
and 8). This LTP was initiated in the first few minutes of activity
and eventually reached a plateau.
We tested a set of four further pairs of place cells (2A and 2B, 2C
and 2D, 3A and 3B, 4A and 4B) with overlapping place fields but with
strikingly different spiking characteristics [for a full description of
the place cell spike pattern characteristics and plasticity outcomes
see Isaac et al. (2009)] and found in each case the model predicted
FIGURE 4 | [Ca2+] dynamics in response to triplets of one pre- and two
post-synaptic spikes. (A) The model calculates [Ca2+] within a spine from
the membrane potential resulting from a triplet of pre- and post-synaptic
spikes. Gray line shows EPSP in the absence of BPAP. (B). Varying t shows
that [Ca2+]max is greatest when 0 t 30 ms for 10 or 20 mV EPSPs for s = 10
ms. (C) Varying s shows that [Ca2+]max decreases as s increases for 10 or
20 mV EPSPs and t = 10 ms. (D) The frequency of spike pairings given at
t = 10 ms and s = 10 ms determines [Ca2+]max.
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 8
Rackham et al. Model for hippocampal synaptic plasticity
post-synaptic then LTD would be induced but the model predicted
only marginal LTD (Figure 10B) in line with the experimental data
(Isaac et al., 2009).
Finally we have compared the experimentally determined plas-
ticity outcome from nine pairs of place cells with the outcome
predicted by our model. We find that the correlation between the
predicted and observed values is significant (Figure 10C, r2 = 0.58,
P < 0.05 by linear regression) and therefore conclude that the model
successfully predicts the induction of synaptic plasticity by irregular
activity patterns.
dIscussIon
The model described in this study incorporates two important
components of Ca2+ dynamics in dendritic spines that are neces-
sary for the induction of synaptic plasticity. Firstly, our model is
capable of analyzing Ca2+ influx and concentration continuously
and therefore it can determine the plasticity outcome of multiple
synaptic events that occur in vivo in an irregular pattern. Secondly,
[Ca2+] is modeled at the synapse in dendritic spines rather than at
the soma. This is important since the critical Ca2+ signal for the
induction of synaptic plasticity occurs at the spine. It also changes
the relative importance of EPSP vs BPAP depolarization which
has major implications for the predicted induction of STDP. This
approach is validated by comparison of the predicted vs observed
Ca2+ transients in response to either a single EPSP or coupled with
a BPAP (Figure 2) (Bloodgood and Sabatini, 2007).
The absolute values for [Ca2+] within the dendrite required for
the induction of synaptic plasticity have been estimated as 150–
500 nM for LTD and >500 nM for LTP (Cormier et al., 2001).
However, other researchers have estimated [Ca2+] within a spine
FIGURE 5 | Theta burst pairing produces large spine [Ca2+]. The model
calculates [Ca2+] within a spine from the membrane potential resulting from
coincident theta burst stimulation of pre- and post-synaptic neurons (black) or
only pre-synaptic neuron (gray).
FIGURE 6 | Post-synaptic voltage clamp paired with pre-synaptic
stimulation determines spine [Ca2+]. The model predicts that voltage clamp
of the post-synaptic membrane potential at 40 mV produces a much smaller
spine [Ca2+] than 0 mV when paired with a single pre-synaptic stimulation.
However, when we reversed the place cell firing patterns such that
cell 1B was pre-synaptic and cell 1A post-synaptic the model pre-
dicted LTP (Figure 10A) that corroborates the experimental results
and closely reproduces the experimentally determined timecourse
of LTP development (Isaac et al., 2009). We also manipulated the
spike patterns in cell 1B to remove all spikes that occurred less than
100 ms after a spike in cell 1A leaving only spikes that occurred
before any spike in cell 1A. Classical STDP rules would again pre-
dict that if cell 1A was pre-synaptic and the modified cell 1B was
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 9
Rackham et al. Model for hippocampal synaptic plasticity
FIGURE 7 | Spine [Ca2+] determines the direction and magnitude of synaptic
weight change. (A). The Ω−function describes the relationship between peak spine
[Ca2+] and synaptic weight change. Symbols represent the peak [Ca2+] produced by a
single application of the plasticity induction protocols shown in Figures 3–6 and
indicate the resulting predicted synaptic weight change. (B) The η-function describes
the learning rate for synaptic weight change as a function of peak spine [Ca2+].
FIGURE 8 | Example of predicted synaptic weight change during overlapping place cell activity. The model calculates spine [Ca2+] during a 16-min period of
activity from two place cells (1A and 1B) with overlapping place fields. The synaptic weight change is then calculated from the peak spine [Ca2+] and shows a robust,
rapidly developing potentiation.
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 10
Rackham et al. Model for hippocampal synaptic plasticity
in response to a single EPSP at 700 nM and a much higher 12 μM
during pairing of post-synaptic depolarization with synaptic stimu-
lation (Sabatini et al., 2002). This discrepancy could be explained
in a number of ways. The [Ca2+] in a dendritic spine in response
to synaptic stimulation could be considerably higher than in the
dendritic shaft because of the diffusion barrier created by the spine
neck. In addition, accurate absolute values for [Ca2+] measured by
fluorescent Ca2+ indicators are difficult to achieve and therefore
most studies are restricted to ratiometric measurements of tran-
sient [Ca2+] increases. For the purposes of synaptic plasticity this
is sufficient since the increase in [Ca2+] triggers induction. Here,
we have calculated the [Ca2+] based on a number of assumptions
for channel conductance and Ca2+ diffusion. More importantly,
we have modeled the relative [Ca2+] increases caused by various
induction protocols and used these to define the graph in Figure 7
that predicts the plasticity outcome.
Inhibitory synaptic transmission has a major role regulating the
induction of synaptic plasticity in the hippocampus. The transient
depression of inhibition induced by activation of pre-synaptic
cannabinoid or GABAB receptors facilitates the induction of LTP
(Davies et al., 1991; Chevaleyre and Castillo, 2004). This modulation
of synaptic plasticity is not included in our current model but incor-
poration of the hyperpolarizing effects of GABAergic transmission
would be an important future improvement and might, for example,
contribute to the frequency dependence of STDP induction.
NMDARs are not the only sources of Ca2+ within dendritic spines
but are certainly the most important for the induction of synaptic
plasticity. A role has also been demonstrated for Ca2+ stores present
in the endoplasmic reticulum in dendrites and spines (linked to Ca2+
influx through NMDARs or mGluRs) and also voltage-dependent
FIGURE 9 | Predicted synaptic weight changes for overlapping and non-overlapping place cell activity. Calculated synaptic weight changes for four pairs of
overlapping place cells 2A, 2B (A), 2C, 2D (B), 3A, 3B (C), and 4A, 4B (D) as well as one pair of non-overlapping place cells 1A, 1C (E) and one pair of adjacent place cells
2E, 2D (F).
FIGURE 10 | Predicted synaptic weight changes for place cell activity with
specific spike patterns. Calculated synaptic weight changes for a pair of
overlapping place cells with an asymmetric cross-correlation 1B, 1A (A) and a pair
of place cells where all spike intervals with positive t less than 100 ms have
been removed 1A, 1B (B). (C) A comparison between the induced plasticity
predicted by the model and the observed plasticity from experimental data
(Isaac et al., 2009).
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 11
Rackham et al. Model for hippocampal synaptic plasticity
Collingridge, G. L., and Isaac, J. T. R.
(2000). PDZ proteins interacting with
C-terminal GluR2/3 are involved in a
PKC-dependent regulation of AMPA
receptors at hippocampal synapses.
Neuron 28, 873–886.
Debanne, D., Gahwiler, B. H., and
Thompson, S. M. (1998). Long-term
synaptic plasticity between pairs of
individual CA3 pyramidal cells in rat
hippocampal slice cultures. J. Physiol.
507, 237–247.
Dickinson, B. A., Jo, J., Seok, H., Son, G.
H., Whitcomb, D. J., Davies, C. H.,
Sheng, M., Collingridge, G. L., and
Cho, K. (2009). A novel mechanism
of hippocampal LTD involving mus-
carinic receptor-triggered interactions
between AMPARs, GRIP and liprin-
alpha. Mol. Brain 2, 18.
Fernandez de Sevilla, D., Fuenzalida,
M., Porto Pazos, A. B., and Buno,
W. (2007). Selective shunting of the
NMDA EPSP component by the slow
afterhyperpolarization in rat CA1
Chevaleyre, V., and Castillo, P. E. (2004).
Endocannabinoid-mediated meta-
plasticity in the hippocampus. Neuron
43, 871–881.
Cho, K., Aggleton, J. P., Brown, M. W., and
Bashir, Z. I. (2001). An experimental
test of the role of postsynaptic cal-
cium levels in determining synaptic
strength using perirhinal cortex of rat.
J. Physiol. 532, 459–466.
Cormier, R. J., Greenwood, A. C., and
Connor, J. A. (2001). Bidirectional
synaptic plasticity correlated with
the magnitude of dendritic calcium
transients above a threshold. J.
Neurophysiol. 85, 399–406.
Dan, Y., and Poo, M. M. (2006). Spike timing-
dependent plasticity: From synapse to
perception. Physiol. Rev. 86, 1033–1048.
Davies, C. H., Starkey, S. J., Pozza, M. F.,
and Collingridge, G. L. (1991). GABA
autoreceptors regulate the induction
of LTP. Nature 349, 609–611.
Daw, M. I., Chittajallu, R., Bortolotto, Z.
A., Dev, K. K., Duprat, F., Henley, J. M.,
ity induction rules and the require-
ment for postsynaptic spikes in rat
hippocampal CA1 pyramidal neu-
rones. J. Physiol. 585, 429–445.
Buchanan, K. A., and Mellor, J. R. (2010).
The activity requirements for spike
timing-dependent plasticity in
the hippocampus. Front. Synaptic
Neurosci. 2, 1–5.
Campanac, E., and Debanne, D. (2008).
Spike timing-dependent plasticity: a
learning rule for dendritic integration
in rat CA1 pyramidal neurons. J.
Physiol. 586, 779–793.
Canepari, M., Djurisic, M., and Zecevic, D.
(2007). Dendritic signals from rat hip-
pocampal CA1 pyr amidal neurons dur-
ing coincident pre- and post-synaptic
activity: a combined voltage- and calcium-
imaging study. J. Physiol. 580, 463–484.
Castellani, G. C., Bazzani, A., and Cooper,
L. N. (2009). Toward a microscopic
model of bidirectional synaptic plas-
ticity. Proc. Natl. Acad. Sci. U.S.A. 106,
14091–14095.
references
Bashir, Z. I., Bortolotto, Z. A., Davies, C.
H., Berretta, N., Irving, A. J., Seal, A. J.,
Henley, J. M., Jane, D. E., Watkins, J. C.,
and Collingridge, G. L. (1993). Induction
of LTP in the hippocampus needs synap-
tic activation of glutamate metabotropic
receptors. Nature 363, 347–350.
Bear, M. F., Cooper, L. N., and Ebner, F.
F. (1987). A physiological-basis for
a theory of synapse modification.
Science 237, 42–48.
Bi, G. Q., and Poo, M. M. (1998). Synaptic
modifications in cultured hippocam-
pal neurons: Dependence on spike
timing, synaptic strength, and post-
synaptic cell type. J. Neurosci. 18,
10464–10472.
Bloodgood, B. L., and Sabatini, B. L. (2007).
Nonlinear regulation of unitary synap-
tic signals by CaV(2.3) voltage-sensitive
calcium channels located in dendritic
spines. Neuron 53, 249–260.
Buchanan, K. A., and Mellor, J. R. (2007).
The development of synaptic plastic-
receptor activation no plasticity is induced but in the presence of
a muscarinic agonist we obtain the results described above. This
suggests that muscarinic receptor activation is permissive for the
induction of synaptic plasticity under conditions where single syn-
apses are activated and EPSP amplitude is small. A similar role has
also been suggested for other G-protein coupled receptors such as
mGluRs (Bashir et al., 1993) and β-adrenoreceptors (Seol et al.,
2007). The observation that muscarinic receptor activation is not
required when multiple synapses are activated and EPSP ampli-
tude is large also demonstrates the critical role that multiple syn-
aptic activation plays in the induction of synaptic plasticity. The
model described in this study does not currently include a role for
G-protein-mediated modulation of synaptic plasticity and this is
an important consideration for future modifications and improve-
ments. For example, it would be potentially possible to incorporate
a negative feedback on Ca2+ influx under normal conditions that
is inhibited by activation of G-protein coupled receptors or over-
ridden by larger EPSPs.
Technical considerations make the visualization or record-
ing of single synapses during behavior extremely difficult. This
makes determination of the events that induce synaptic plasticity
in the hippocampus almost impossible. These limitations could
be circumvented by using the model described in this study to
predict the plasticity outcome at a single synapse for any pattern
of pre- and post-synaptic activity. Potentially, this will enable the
determination of when synaptic plasticity is induced during behav-
ior by examining the spike patterns of pairs of cells recorded during
a memory task. This represents an exciting future application for
this model.
acknowledgMents
Jack R. Mellor is supported by the Wellcome Trust and by the
European Union ENI-NET. The Bristol Centre for Complexity
Sciences is funded by the EPSRC, UK (EP/E501214/1) which sup-
ports OJLR and AG.
calcium channels present on the membrane of dendritic spines that
generate local dendritic Ca2+ spikes (Golding et al., 2002; Remy and
Spruston, 2007) and can contribute to some forms of LTD (Nevian
and Sakmann, 2006). The contribution of non-NMDAR-mediated
Ca2+ increases to synaptic plasticity induction is specifically omitted
from our model but would be a useful future addition. There are
also potential alternative mechanisms for the induction of synaptic
plasticity that do not involve Ca2+ such as mGluRs (Moult et al.,
2006) or muscarinic acetylcholine receptors (Dickinson et al., 2009)
that are also specifically omitted from our model. Future experi-
mental data may, however, require their addition.
There is a wealth of data on the spiking requirements for plastic-
ity at various synapses in the cortex (Dan and Poo, 2006) . H owe ver,
these requirements have been shown to be somewhat different to
those found in the hippocampus perhaps due to variations in den-
dritic architecture and NMDAR subunit expression and therefore we
have restricted our model specifically to the Schaffer collateral syn-
apse in the hippocampus. Even within the hippocampus there is some
controversy surrounding the precise spiking requirements for STDP
(Buchanan and Mellor, 2010) for example whether single pairs of spikes
at low frequency can induce LTP (Bi and Poo, 1998; Debanne et al.,
1998; Nishiyama et al., 2000; Wittenberg and Wang, 2006; Buchanan
and Mellor, 2007; Campanac and Debanne, 2008; Kwag and Paulsen,
2009). These discrepancies can potentially be explained by variations
in experimental conditions [For example, size of EPSPs, age of animals,
frequency of stimulus presentation, presence of inhibitory inputs, etc.
discussed in Buchanan and Mellor (2010)]. To reduce these discrep-
ancies our model is based on data from acute hippocampal slices but
future modifications could enable it to accurately replicate data from
all systems including cortical synapses if one assumes that the general
mechanism for the induction of synaptic plasticity is similar for all
synapses that have NMDAR-dependent synaptic plasticity.
Our experimental data has revealed a critical role for muscarinic
acetylcholine receptors in facilitating LTP induced by place cell
firing patterns (Isaac et al., 2009). In the absence of muscarinic
Frontiers in Synaptic Neuroscience www.frontiersin.org July 2010 | Volume 2 | Article 31 | 12
Rackham et al. Model for hippocampal synaptic plasticity
pyramidal neurons. J. Neurophysiol.
97, 3242–3255.
Franks, K. M., Bartol, T. M. Jr., and
Sejnowski, T. J. (2002). A Monte Carlo
model reveals independent signaling
at central glutamatergic synapses.
Biophys. J. 83, 2333–2348.
Frick, A., Magee, J., and Johnston, D. (2004).
LTP is accompanied by an enhanced
local excitability of pyramidal neuron
dendrites. Nat. Neurosci. 7, 126–135.
Froemke, R. C., Tsay, I. A., Raad, M., Long,
J. D., and Dan, Y. (2006). Contribution
of individual spikes in burst-induced
long-term synaptic modification. J.
Neurophysiol. 95, 1620–1629.
Gall, D., Prestori, F., Sola, E., D’Errico,
A., Roussel, C., Forti, L., Rossi, P., and
D’Angelo, E. (2005). Intracellular cal-
cium regulation by burst discharge
determines bidirectional long-term syn-
aptic plasticity at the cerebellum input
stage. J. Neurosci. 25, 4813–4822.
Golding, N. L., Staff, N. P., and Spruston,
N. (2002). Dendritic spikes as a mecha-
nism for cooperative long-term poten-
tiation. Nature 418, 326–331.
Graupner, M., and Brunel, N. (2007).
STDP in a bistable synapse model
based on CaMKII and associated sig-
naling pathways. PLoS Comput. Biol.
3, 2299–2323.
Grunditz, A., Holbro, N., Tian, L., Zuo,
Y., and Oertner, T. G. (2008). Spine
neck plasticity controls postsynaptic
calcium signals through electrical
compartmentalization. J. Neurosci.
28, 13457–13466.
Hansel, C., Artola, A., and Singer, W.
(1996). Different threshold levels of
postsynaptic [Ca2+]i have to be reached
to induce LTP and LTD in neocorti-
cal pyramidal cells. J. Physiol. Paris 90,
317–319.
Hansel, C., Ar tola, A., and Singer, W. (1997).
Relation between dendritic Ca2+ levels
and the polarity of synaptic long-term
modifications in rat visual cortex neu-
rons. Eur. J. Neurosci. 9, 2309–2322.
Helias, M., Rotter, S., Gewaltig, M. O.,
and Diesmann, M. (2008). Structural
plasticity controlled by calcium based
correlation detection. Front. Comput.
Neurosci. 2, 1–21.
Isaac, J. T., Buchanan, K. A., Muller, R. U.,
and Mellor, J. R. (2009). Hippocampal
place cell firing patterns can induce
long-term synaptic plasticity in vitro.
J. Neurosci. 29, 6840–6850.
Isaac, J. T., Hjelmstad, G. O., Nicoll, R. A.,
and Malenka, R. C. (1996). Long-term
potentiation at single fiber inputs to
hippocampal CA1 pyramidal cells.
Proc. Natl. Acad. Sci. U.S.A. 93,
8710–8715.
Isaac, J. T., Nicoll, R. A., and Malenka, R.
C. (1995). Evidence for silent synapses:
implications for the expression of LTP.
Neuron 15, 427–434.
Ismailov, I., Kalikulov, D., Inoue, T., and
Friedlander, M. J. (2004). The kinetic
profile of intracellular calcium pre-
dicts long-term potentiation and
long-term depression. J. Neurosci. 24,
9847–9861.
Jahr, C. E., and Stevens, C. F. (1990). A
quantitative description of NMDA
receptor-channel kinetic behavior. J.
Neurosci. 10, 1830–1837.
Keller, D. X., Franks, K. M., Bartol, T.
M. Jr., and Sejnowski, T. J. (2008).
Calmodulin activation by calcium
transients in the postsynaptic den-
sity of dendritic spines. PLoS One 3,
e2045.
Kwag, J., and Paulsen, O. (2009). The
timing of external input controls the
sign of plasticity at local synapses. Nat.
Neurosci. 12, 1219–1221.
Larson, J., Wong, D., and Lynch, G. (1986).
Patterned stimulation at the theta-fre-
quency is optimal for the induction of
hippocampal long-term potentiation.
Brain Res. 368, 347–350.
Lisman, J. (1989). A mechanism for the
Hebb and the anti-Hebb processes
underlying learning and memory.
Proc. Natl. Acad. Sci. U.S.A. 86,
9574–9578.
Miller, P., Zhabotinsky, A. M., Lisman, J.
E., and Wang, X. J. (2005). The sta-
bility of a stochastic CaMKII switch:
dependence on the number of enzyme
molecules and protein turnover. PLoS
Biol. 3, e107. doi:10.1371/journal.
pbio.0030107.
Moult, P. R., Gladding, C. M., Sanderson,
T. M., Fitzjohn, S. M., Bashir, Z. I.,
Molnar, E., and Collingridge, G. L.
(2006). Tyrosine phosphatases regu-
late AMPA receptor trafficking during
metabotropic glutamate receptor-
mediated long-term depression. J.
Neurosci. 26, 2544–2554.
Nevian, T., and Sakmann, B. (2006).
Spine Ca2+ signaling in spike-timing-
dependent plasticity. J. Neurosci. 26,
11001–11013.
Nishiyama, M., Hong, K., Mikoshiba, K.,
Poo, M., and Kato, K. (2000). Calcium
stores regulate the polarity and input
specificity of synaptic modification.
Nature 408, 584–588.
Palmer, L. M., and Stuart, G. J. (2009).
Membrane potential changes in den-
dritic spines during action potentials
and synaptic input. J. Neurosci. 29,
6897–6903.
Pike, F. G., Meredith, R. M., Olding,
A. W. A., and Paulsen, O. (1999).
Postsynaptic bursting is essential for
‘Hebbian’ induction of associative
long-term potentiation at excita-
tory synapses in rat hippocampus. J.
Physiol. 518, 571–576.
Remy, S., and Spruston, N. (2007).
Dendritic spikes induce single-burst
long-term potentiation. Proc. Natl.
Acad. Sci. U.S.A. 104, 17192–17197.
Rubin, J. E., Gerkin, R. C., Bi, G. Q., and
Chow, C. C. (2005). Calcium time
course as a signal for spike-timing-
dependent plasticity. J. Neurophysiol.
93, 2600–2613.
Sabatini, B. L., Oertner, T. G., and
Svoboda, K. (2002). The life cycle of
Ca2+ ions in dendritic spines. Neuron
33, 439–452.
Seol, G. H., Ziburkus, J., Huang, S., Song,
L., Kim, I. T., Takamiya, K., Huganir,
R. L., Lee, H. K., and Kirkwood, A.
(2007). Neuromodulators control
the polarity of spike-timing-de-
pendent synaptic plasticity. Neuron
55, 919–929.
Shouval, H. Z., Bear, M. F., and Cooper, L.
N. (2002). A unified model of NMDA
receptor-dependent bidirectional syn-
aptic plasticity. Proc. Natl. Acad. Sci.
U.S.A. 99, 10831–10836.
Sjostrom, P. J., and Nelson, S. B. (2002).
Spike timing, calcium signals and
synaptic plasticity. Curr. Opin.
Neurobiol. 12, 305–314.
Sjostrom, P. J., Turrigiano, G. G., and
Nelson, S. B. (2001). Rate, timing,
and cooperativity jointly determine
cortical synaptic plasticity. Neuron
32, 1149–1164.
Sobczyk, A., and Svoboda, K. (2007).
Activity-dependent plasticity of the
NMDA-receptor fractional Ca2+ cur-
rent. Neuron 53, 17–24.
Song, S., Miller, K. D., and Abbott, L. F.
(2000). Competitive Hebbian learn-
ing through spike-timing-dependent
synaptic plasticity. Nat. Neurosci. 3,
919–926.
Urakubo, H., Honda, M., Froemke, R. C.,
and Kuroda, S. (2008). Requirement
of an allosteric kinetics of NMDA
receptors for spike timing-de-
pendent plasticity. J. Neurosci. 28,
3310–3323.
Wittenberg, G. M., and Wang, S. S. H.
(2006). Malleability of spike-timing-
dependent plasticity at the CA3-CA1
synapse. J. Neurosci. 26, 6610–6617.
Conflict of Interest Statement: The
authors declare that the research was
conducted in the absence of any com-
mercial or financial relationships that
could be construed as a potential conflict
of interest.
Received: 21 January 2010; paper pend-
ing published: 22 February 2010; accepted:
27 June 2010; published online: 21 July
2010.
Citation: Rackham OJL, Tsaneva-
Atanasova K, Ganesh A and Mellor JR
(2010) A Ca2+-based computational model
for NMDA receptor-dependent synaptic
plasticity at individual post-synaptic spines
in the hippocampus. Front. Syn. Neurosci.
2:31. doi: 10.3389/fnsyn.2010.00031
Copyright © 2010 Rackham, Tsaneva-
Atanasova, Ganesh and Mellor. This is an
open-access article subject to an exclusive
license agreement between the authors
and the Frontiers Research Foundation,
which permits unrestricted use, distribu-
tion, and reproduction in any medium,
provided the original authors and source
are credited.
... where V max is the maximum depolarization value for bAP value set to +67 mV [58], I fast and I slow are the relative magnitudes of the fast and slow components of the bAP that sum to one, and τ fast and τ slow are the relative time constants that describe the exponential decays of the two components. ...
... Many synaptic models that have been proposed in the past estimated the EPSCs and/or the EPSPs simply by using two-component exponential functions fitted on electrophysiological recordings [35,[58][59][60]. On the contrary, in our model, the open probabilities of the receptors vary according to a system of deterministic rate equations that represent massaction kinetics of receptors-neurotransmitter interactions [53]. ...
... where G Ca 2+ expresses the permeability of the NMDA receptor to Ca 2+ ions, set to 2 nM·ms −1 ·mV −1 [58] and V E Ca 2+ is the reversal equilibrium potential for Ca 2+ set to +130 mV [58]. Finally, the calcium dynamics in the post-synaptic cell is integrated by a simple first-order differential equation [35,58]: ...
Article
Full-text available
Inside hippocampal circuits, neuroplasticity events that individual cells may undergo during synaptic transmissions occur in the form of Long-Term Potentiation (LTP) and Long-Term Depression (LTD). The high density of NMDA receptors expressed on the surface of the dendritic CA1 spines confers to hippocampal CA3-CA1 synapses the ability to easily undergo NMDA-mediated LTP and LTD, which is essential for some forms of explicit learning in mammals. Providing a comprehensive kinetic model that can be used for running computer simulations of the synaptic transmission process is currently a major challenge. Here, we propose a compartmentalized kinetic model for CA3-CA1 synaptic transmission. Our major goal was to tune our model in order to predict the functional impact caused by disease associated variants of NMDA receptors related to severe cognitive impairment. Indeed, for variants Glu413Gly and Cys461Phe, our model predicts negative shifts in the glutamate affinity and changes in the kinetic behavior, consistent with experimental data. These results point to the predictive power of this multiscale viewpoint, which aims to integrate the quantitative kinetic description of large interaction networks typical of system biology approaches with a focus on the quality of a few, key, molecular interactions typical of structural biology ones.
... However, experimental data suggests that calcium dynamics are also important (Yang et al., 1999;Mizuno et al., 2001;Wang et al., 2005;Nevian and Sakmann, 2006;Tigaret et al., 2016). As a result, subsequent phenomenological models of plasticity incorporated slow variables that integrate the fast synaptic input signals, loosely modelling calcium and its downstream effectors (Abarbanel et al., 2003;Rubin et al., 2005;Rackham et al., 2010;Kumar and Mehta, 2011;Graupner and Brunel, 2012;Honda et al., 2013;Standage et al., 2014;De Pittà and Brunel, 2016). Concurrently, more detailed models tried to explicitly describe the molecular pathways integrating the calcium dynamics and its stochastic nature (Cai et al., 2007;Shouval and Kalantzis, 2005;Miller et al., 2005;Zeng and Holmes, 2010;Yeung et al., 2004). ...
... In contrast, neural spiking in hippocampus in vivo is irregular and variable (Fenton and Muller, 1998;Isaac et al., 2009). Previous studies that asked how natural firing variability affects the rules of plasticity induction used simpler synapse models (Rackham et al., 2010;Graupner et al., 2016;Cui et al., 2018). We explored this question in our synapse model using simulations with three distinct types of additional variability: (1) spike time jitter, (2) failures induced by dropping spikes, (3) independent pre and postsynaptic Poisson spike trains (Graupner et al., 2016). ...
Article
Full-text available
Discovering the rules of synaptic plasticity is an important step for understanding brain learning. Existing plasticity models are either (1) top-down and interpretable, but not flexible enough to account for experimental data, or (2) bottom-up and biologically realistic, but too intricate to interpret and hard to fit to data. To avoid the shortcomings of these approaches, we present a new plasticity rule based on a geometrical readout mechanism that flexibly maps synaptic enzyme dynamics to predict plasticity outcomes. We apply this readout to a multi-timescale model of hippocampal synaptic plasticity induction that includes electrical dynamics, calcium, CaMKII and calcineurin, and accurate representation of intrinsic noise sources. Using a single set of model parameters, we demonstrate the robustness of this plasticity rule by reproducing nine published ex vivo experiments covering various spike-timing and frequency-dependent plasticity induction protocols, animal ages, and experimental conditions. Our model also predicts that in vivo-like spike timing irregularity strongly shapes plasticity outcome. This geometrical readout modelling approach can be readily applied to other excitatory or inhibitory synapses to discover their synaptic plasticity rules.
... Based on these findings [29] had designed a model of plasticity in spines that predicts that an EPSP resulting from the activation of a single synapse is sufficient to cause a significant Ca ++ influx through NMDA receptors. This is in line with experimental data [30][31][32]. ...
Article
Full-text available
In many situations it is behaviorally relevant for an animal to respond to co-occurrences of perceptual, possibly polymodal features, while these features alone may have no importance. Thus, it is crucial for animals to learn such feature combinations in spite of the fact that they may occur with variable intensity and occurrence frequency. Here, we present a novel unsupervised learning mechanism that is largely independent of these contingencies and allows neurons in a network to achieve specificity for different feature combinations. This is achieved by a novel correlation-based (Hebbian) learning rule, which allows for linear weight growth and which is combined with a mechanism for gradually reducing the learning rate as soon as the neuron’s response becomes feature combination specific. In a set of control experiments, we show that other existing advanced learning rules cannot satisfactorily form ordered multi-feature representations. In addition, we show that networks, which use this type of learning always stabilize and converge to subsets of neurons with different feature-combination specificity. Neurons with this property may, thus, serve as an initial stage for the processing of ecologically relevant real world situations for an animal.
... These types of STDP models vary in their consideration of the postsynaptic [Ca 2+ ] transient, with some using peak magnitude only as their signal (e.g., Shouval et al. 2002; (Lisman 1989). Examples of models using this approach include Shouval et al. (2002), Abarbanel et al. (2003), Rubin et al. (2005), and Rackham et al. (2010). The simplest of these models (Shouval et al. 2002, Fig. 1) results in a second LTD window at large positive STDP time intervals between pre-and postsynaptic spikes, an effect that is not observed in the majority of experimental STDP spike-pair studies. ...
... The synaptic models used are biophysically motivated and account for relevant physiological observations. Due to their voltage-gated component, the dynamics of NMDA receptors (NMDRs) mediates the generation of sustained plateau potentials (Major et al., 2008) and supports coincidence detection (Tabone & Ramaswami, 2012;Rackham et al., 2010). It affects the integration of excitatory input in the dendrites and the soma, and plays a key role in shaping dendritic processing, synaptic plasticity, and the global input-output behavior of neurons (Doron et al., 2017;Smith et al., 2013). ...
Article
Full-text available
Key points: We present a neuron model, called the Tripod, with two segregated dendritic branches that are connected to the axosomatic compartment. Each branch implements inhibitory GABAergic and excitatory glutamatergic synaptic transmission, including voltage-gated NMDA receptors. Dendrites are modeled on relevant geometric and physiological parameters measured in human pyramidal cells. The neuron reproduces classical dendritic computations, such as coincidence detection and pathway selection via shunting inhibition, beyond the scope of point-neuron models. Under some conditions, dendritic NMDA spikes cause plateau potentials and we show that they provide a form of short-term memory which is useful for sequence recognition. The dendritic structure of the Tripod neuron is sufficiently simple to be integrated into efficient network simulations and studied in a broad functional context. Abstract: Neuron models with explicit dendritic dynamics have shed light on mechanisms for coincidence detection, pathway selection, and temporal filtering. However, it is still unclear which morphological and physiological features are required to capture these phenomena. In this work, we introduce the Tripod neuron model and propose a minimal structural reduction of the dendritic tree that is able to reproduce these dendritic computations. The Tripod is a three-compartment model consisting of two segregated passive dendrites and a somatic compartment modeled as an adaptive, exponential integrate-and-fire neuron. It incorporates dendritic geometry, membrane physiology, and receptor dynamics as measured in human pyramidal cells. We characterize the response of the Tripod to glutamatergic and GABAergic inputs and identify parameters that support supra-linear integration, coincidence-detection, and pathway-specific gating through shunting inhibition. Following NMDA spikes, the Tripod neuron generates plateau potentials whose duration depends on the dendritic length and the strength of synaptic input. When fitted with distal compartments, the Tripod neuron encodes previous activity into a dendritic depolarized state. This dendritic memory allows the neuron to perform temporal binding and we show that the neuron solves transition and sequence detection tasks on which a single-compartment model fails. Thus, the Tripod neuron can account for dendritic computations previously explained only with more detailed neuron models or neural networks. Due to its simplicity, the Tripod model can be used efficiently in simulations of larger cortical circuits. Abstract figure legend Cardiopulmonary and cerebrovascular function continues to develop through childhood and adolescence, but the impact of maturation on acclimatization responses to high-altitude is unknown. 10 children (9.8±2.5 y) and 10 adults (34.7±7.1 y) were passively brought from sea level to high-altitude and measurements of minute ventilation (VE), stroke volume (SV), pulmonary artery systolic pressure (PASP), and cerebral blood flow (CBF) were obtained at 3000 m and twice at 3800 m. Children demonstrated a greater increase in VE, a smaller increase in PASP, and a greater reduction in SV, paired with larger chronotropic response at 3000 m compared to adults, but age-related differences in cardiopulmonary function and ventilatory changes diminished at 3800 m. Baseline CBF was consistently elevated in children compared to adults, but relative changes in CBF with altitude were similar between children and adults at all times.
... The synaptic models used are biophysically motivated and account for relevant physiological observations. Due to their voltage-gated component, the dynamics of NMDA receptors (NMDRs) mediates the generation of sustained plateau potentials (Major et al., 2008) and supports coincidence detection (Tabone & Ramaswami, 2012;Rackham et al., 2010). It affects the integration of excitatory input in the dendrites and the soma, and plays a key role in shaping dendritic processing, synaptic plasticity, and the global input-output behavior of neurons (Doron et al., 2017;Smith et al., 2013). ...
Preprint
Full-text available
Neuron models with explicit dendritic dynamics have shed light on mechanisms for coincidence detection, pathway selection, and temporal filtering. However, it is still unclear which morphological and physiological features are required to capture these phenomena. In this work, we introduce the Tripod neuron model and propose a minimal structural reduction of the dendritic tree that is able to reproduce these dendritic computations. The Tripod is a three-compartment model consisting of two segregated passive dendrites and a somatic compartment modeled as an adaptive, exponential integrate-and-fire neuron. It incorporates dendritic geometry, membrane physiology, and receptor dynamics as measured in human pyramidal cells. We characterize the response of the Tripod to glutamatergic and GABAergic inputs and identify parameters that support supra-linear integration, coincidence-detection, and pathway-specific gating through shunting inhibition. Following NMDA spikes, the Tripod neuron generates plateau potentials whose duration depends on the dendritic length and the strength of synaptic input. When fitted with distal compartments, the Tripod neuron encodes previous activity into a dendritic depolarized state. This dendritic memory allows the neuron to perform temporal binding and we show that the neuron solves transition and sequence detection tasks on which a single-compartment model fails. Thus, the Tripod neuron can account for dendritic computations previously explained only with more detailed neuron models or neural networks. Due to its simplicity, the Tripod model can be used efficiently in simulations of larger cortical circuits.
... The main methodological development that enabled our analysis is the construction of the stochastic model that, unlike the models based on average reaction-diffusion equations, which usually ignore molecular details [46,114,130,133,142,152,163,165,166], integrates local binding, and organization of molecules, channels and transporters [167,168]. Therefore, the stochastic modeling allows monitoring of diffusion trajectory of each ion separately, thus providing insights into the level of single molecules usually at the expense of high computational cost. The ability to analyze a system at the single molecule (stochastic) manner is especially important for a system like dendritic spines, given that many of spine processes occur at low copy numbers. ...
Thesis
Dendritic spines that form the receiving terminals of neuronal signals are regulated biochemically by calcium. Fast calcium transmissions in spines remain poorly understood as they occur in a few milliseconds and cannot be explained by classical diffusion theory. I developed mathematical models and stochastic simulations based on extreme statistics to compute the time taken by the fastest ions to arrive at a target via an optimal path. This novel framework not only explained fast calcium transients in spines, but also predicted accurately the spatial distributions of three calcium channels in spines controlling intake, storage and release. I also modeled their interplay during calcium regulation that could lead to synaptic changes underlying learning and memory. These new paradigms of nanoscale molecular organization and extreme statistics could also characterize the timescales of many other biophysical processes driven by random arrivals of the fastest particles to a small target.
... Many studies have tried to establish if calcium alone can account for synaptic plasticity (Shouval et al., 2002;Rackham et al., 2010;Chindemi et al., 2020), but given the involvement of multiple mechanisms based on dendrite biophysics (Froemke et al., 2010), presynaptic NMDA receptors (Wong et al., 2021) or NMDA receptors that signal unconventionally (Dore et al., 2017), metabotropic glutamate receptors (Nevian and Sakmann, 2006), and learning mechanisms residing in a third cell such as astrocytes (Min and Nevian, 2012;Adamsky et al., 2018), can we really hope to find one calcium model that fits all synapse types? If the calcium hypothesis in synaptic plasticity is not valid for all synapse types, then can we at least establish a phenomenological model that captures plasticity at all synapses? ...
... The main methodological development that enabled our analysis is the construction of the stochastic model that, unlike the models based on average reaction-diffusion equations, which usually ignore molecular details (25,(29)(30)(31)(32)(33)(34)(35)(36), integrates local binding and organization of molecules, channels, and transporters (37,38). Therefore, the stochastic modeling allows monitoring of diffusion trajectory of each ion separately, thus providing insights into the level of single molecules usually at the expense of high computational cost. ...
Article
Full-text available
Dendritic spines are critical components of neuronal synapses as they receive and transform synaptic inputs into a succession of calcium-regulated biochemical events. The spine apparatus (SA), an extension of smooth endoplasmic reticulum, regulates slow and fast calcium dynamics in spines. Calcium release events deplete SA calcium ion reservoir rapidly, yet the next cycle of signaling requires its replenishment. How spines achieve this replenishment without triggering calcium release remains unclear. Using computational modeling, calcium and STED superresolution imaging, we show that the SA replenishment involves the store-operated calcium entry pathway during spontaneous calcium transients. We identified two main conditions for SA replenishment without depletion: a small amplitude and a slow timescale for calcium influx, and a close proximity between SA and plasma membranes. Thereby, spine’s nanoscale organization separates SA replenishment from depletion. We further conclude that spine’s receptor organization also determines the calcium dynamics during the induction of long-term synaptic changes.
Thesis
Full-text available
How learning occurs has been a long-standing question in neuroscience. Since the first demonstration that the strength wiring up neurons can be persistent, the study of neuronal connections, the synapses, became a path to understanding memory formation. In the 70s, the first electrophysiology methods to modify the synaptic strength were discovered, leading to the evidence of how synapses subjected to stimulation are plastic. Such a form of synaptic plasticity was predicted two decades before by a theoretical synaptic rule coined by the neuropsychologist Donald Hebb. The possibility that a devised rule could explain memory motivated the birth of theories providing new questions and mechanistic representations of the brain's functioning. The diversification of techniques allowed researchers to investigate in depth the nature of synaptic rules. However, the heterogeneity of experimental conditions adopted by different laboratories implicated that the same stimulation pattern could produce different synaptic modifications. The observed heterogeneity in the methods and outcomes have hindered the formalization of a coherent view on how synaptic plasticity works. To fill this gap, this thesis developed a stochastic computational model of the rat CA3-CA1 glutamatergic synapse to explain and gain insights into how experimental conditions affect plasticity outcomes. I uncovered a new plasticity rule that accounts for methodological differences such as developmental aspects, extracellular medium and temperature influences on synaptic plasticity outcome. The model relies on an expanded version of the previous methods to predict synaptic plasticity, modified to handle combined dynamics. That is achieved by introducing a geometrical readout to interpret the dynamics of two calcium-binding enzymes controlling the induction of plasticity. In this way, the model covers classical and recent stimulation paradigms (e.g. STDP, FDP) using a single rule parameter set. Finally, the model's robustness is tested for in vivo-like spike time irregularity showing how different protocols converge to the same outcome when regularity is altered. This model allows one to obtain testable predictions since it links the simulated variables to the specificity needed to describe a plasticity protocol. Although the model is specific to a single CA3-CA1 synapse, the study's insights may be generalized to other types, enabling a deeper understanding of the rules of synaptic plasticity and learning.
Article
Full-text available
The calcium/calmodulin-dependent protein kinase II (CaMKII) plays a key role in the induction of long-term postsynaptic modifications following calcium entry. Experiments suggest that these long-term synaptic changes are all-or-none switch-like events between discrete states. The biochemical network involving CaMKII and its regulating protein signaling cascade has been hypothesized to durably maintain the evoked synaptic state in the form of a bistable switch. However, it is still unclear whether experimental LTP/LTD protocols lead to corresponding transitions between the two states in realistic models of such a network. We present a detailed biochemical model of the CaMKII autophosphorylation and the protein signaling cascade governing the CaMKII dephosphorylation. As previously shown, two stable states of the CaMKII phosphorylation level exist at resting intracellular calcium concentration, and high calcium transients can switch the system from the weakly phosphorylated (DOWN) to the highly phosphorylated (UP) state of the CaMKII (similar to a LTP event). We show here that increased CaMKII dephosphorylation activity at intermediate Ca2+ concentrations can lead to switching from the UP to the DOWN state (similar to a LTD event). This can be achieved if protein phosphatase activity promoting CaMKII dephosphorylation activates at lower Ca2+ levels than kinase activity. Finally, it is shown that the CaMKII system can qualitatively reproduce results of plasticity outcomes in response to spike-timing dependent plasticity (STDP) and presynaptic stimulation protocols. This shows that the CaMKII protein network can account for both induction, through LTP/LTD-like transitions, and storage, due to its bistability, of synaptic changes.
Article
Full-text available
Synaptic plasticity has historically been investigated most intensely in the hippocampus and therefore it is somewhat surprising that the majority of studies on spike timing-dependent plasticity (STDP) have focused not in the hippocampus but on synapses in the cortex. One of the major reasons for this bias is the relative ease in obtaining paired electrophysiological recordings from synaptically coupled neurons in cortical slices, in comparison to hippocampal slices. Another less obvious reason has been the difficulty in achieving reliable STDP in the hippocampal slice preparation and confusion surrounding the conditions required. The original descriptions of STDP in the hippocampus was performed on paired recordings from neurons in dissociated or slice cultures utilizing single pairs of presynaptic and postsynaptic spikes and were subsequently replicated in acute hippocampal slices. Further work in several laboratories using conditions that more closely replicate the situation in vivo revealed a requirement for multiple postsynaptic spikes that necessarily complicate the absolute timing rules for STDP. Here we review the hippocampal STDP literature focusing on data from acute hippocampal slice preparations and highlighting apparently contradictory results and the variations in experimental conditions that might account for the discrepancies. We conclude by relating the majority of the available experimental data to a model for STDP induction in the hippocampus based on a critical role for postsynaptic Ca(2+) dynamics.
Article
Full-text available
The method by which local networks in the brain store information from extrinsic afferent inputs is not well understood. We found that the timing of afferent input can bidirectionally control the sign of spike timing-dependent plasticity at local synapses in rat hippocampus. This mechanism provides a means by which temporal information in external input can be encoded in the local matrix of synaptic weights.
Article
• The biologically relevant rules of synaptic potentiation were investigated in hippocampal slices from adult rat by mimicking neuronal activity seen during learning behaviours. Synaptic efficacy was monitored in two separate afferent pathways among the Schaffer collaterals during intracellular recording of CA1 pyramidal neurones. The effects of pairing presynaptic single spikes or bursts with postsynaptic single spikes or bursts, repeated at 5 Hz (‘theta’ frequency), were compared. • The pairing of ten single evoked excitatory synaptic events with ten postsynaptic single action potentials at 5 Hz, repeated twelve times, failed to induce synaptic enhancement (EPSP amplitude 95 % of baseline amplitude 20 min after pairing; n= 5 ). In contrast, pairing the same number of action potentials, but clustered in bursts, induced robust synaptic potentiation (EPSP amplitude 163 %; P t test; n= 5 ). This potentiation was input specific, long lasting (> 1 h; n= 3 ) and its induction was blocked by an antagonist at NMDA receptors (20-50 M D(-)-2-amino-5-phosphonopentanoic acid; EPSP amplitude 109 %; n= 6 ). • Presynaptic bursting paired with postsynaptic single action potentials did not induce input specific synaptic change (113 % in the test input vs. 111 % in the control; n= 8 ). In contrast, postsynaptic bursting when paired with presynaptic single action potentials was sufficient to induce synaptic potentiation when the presynaptic activity preceded the postsynaptic activity by 10 ms (150 vs. 84 % in the control input; P n= 10 ). • These results indicate that, under our conditions, postsynaptic bursting activity is necessary for associative synaptic potentiation at CA1 excitatory synapses in adult hippocampus. The existence of a distinct postsynaptic signal for induction of synaptic change calls for refinement of the common interpretation of Hebb's rule, and is likely to have important implications for our understanding of cortical network operation.
Article
Long‐term potentiation (LTP) and depression (LTD) were investigated at synapses formed by pairs of monosynaptically connected CA3 pyramidal cells in rat hippocampal slice cultures. An N ‐methyl‐D‐aspartate (NMDA) receptor‐mediated component of the unitary EPSP, elicited at the resting membrane potential in response to single action potentials in an individual CA3 cell, could be isolated pharmacologically. Associative LTP was induced when single presynaptic action potentials were repeatedly paired with 240 ms postsynaptic depolarizing pulses that evoked five to twelve action potentials or with single postsynaptic action potentials evoked near the peak of the unitary EPSP. LTP induction was prevented by an NMDA receptor antagonist. Associative LTD was induced when single presynaptic action potentials were repeatedly elicited with a certain delay after either 240 ms postsynaptic depolarizing pulses or single postsynaptic action potentials. The time window within which presynaptic activity had to occur for LTD induction was dependent on the amount of postsynaptic depolarization. LTD was induced if single pre‐ and postsynaptic action potentials occurred synchronously. Homosynaptic LTD was induced by 3 Hz tetanization of the presynaptic neuron for 3 min and was blocked by an NMDA receptor antagonist. Depotentiation was produced with stimulation protocols that elicit either homosynaptic or associative LTD. Recurrent excitatory synapses between CA3 cells display associative potentiation and depression. The sign of the change in synaptic strength is a function of the relative timing of pre‐ and postsynaptic action potentials.
Article
Olfactory perception is initiated by the recognition of odorants by a large repertoire of receptors in the sensory epithelium. A dispersed pattern of neural activity in the nose is converted into a segregated map in the olfactory bulb. How is this representation transformed at the next processing center for olfactory information, the piriform cortex? Optical imaging of odorant responses in the cortex reveals that the piriform discards spatial segregation as well as chemotopy and returns to a highly distributed organization in which different odorants activate unique but dispersed ensembles of cortical neurons. Neurons in piriform cortex, responsive to a given odorant, are not only distributed without apparent spatial preference but exhibit discontinuous receptive fields. This representation suggests organizational principles that differ from those in neocortical sensory areas where cells responsive to similar stimulus features are clustered and response properties vary smoothly across the cortex.
Article
We show that a 2-step phospho/dephosphorylation cycle for the alpha-amino-3-hydroxy-5-methyl-4-isoxazole proprionic acid receptor (AMPAR), as used in in vivo learning experiments to assess long-term potentiation (LTP) induction and establishment, exhibits bistability for a wide range of parameters, consistent with values derived from biological literature. The AMPAR model we propose, hence, is a candidate for memory storage and switching behavior at a molecular-microscopic level. Furthermore, the stochastic formulation of the deterministic model leads to a mesoscopic interpretation by considering the effect of enzymatic fluctuations on the Michelis-Menten average dynamics. Under suitable hypotheses, this leads to a stochastic dynamical system with multiplicative noise whose probability density evolves according to a Fokker-Planck equation in the Stratonovich sense. In this approach, the probability density associated with each AMPAR phosphorylation state allows one to compute the probability of any concentration value, whereas the Michaelis-Menten equations consider the average concentration dynamics. We show that bistable dynamics are robust for multiplicative stochastic perturbations and that the presence of both noise and bistability simulates LTP and long-term depression (LTD) behavior. Interestingly, the LTP part of this model has been experimentally verified as a result of in vivo, one-trial inhibitory avoidance learning protocol in rats, that produced the same changes in hippocampal AMPARs phosphorylation state as observed with in vitro induction of LTP with high-frequency stimulation (HFS). A consequence of this model is the possibility of characterizing a molecular switch with a defined biochemical set of reactions showing bistability and bidirectionality. Thus, this 3-enzymes-based biophysical model can predict LTP as well as LTD and their transition rates. The theoretical results can be, in principle, validated by in vitro and in vivo experiments, such as fluorescence measurements and electrophysiological recordings at multiple scales, from molecules to neurons. A further consequence is that the bistable regime occurs only within certain parametric windows, which may simulate a "history-dependent threshold". This effect might be related to the Bienenstock-Cooper-Munro theory of synaptic plasticity.