Download Glycine Binding Sites of Presynaptic NMDA Receptors May

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

Apical dendrite wikipedia , lookup

Environmental enrichment wikipedia , lookup

Single-unit recording wikipedia , lookup

Neuroesthetics wikipedia , lookup

Neuroanatomy wikipedia , lookup

Development of the nervous system wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Optogenetics wikipedia , lookup

Biological neuron model wikipedia , lookup

Signal transduction wikipedia , lookup

Aging brain wikipedia , lookup

Nonsynaptic plasticity wikipedia , lookup

Binding problem wikipedia , lookup

Channelrhodopsin wikipedia , lookup

End-plate potential wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Nervous system network models wikipedia , lookup

Spike-and-wave wikipedia , lookup

Pre-Bötzinger complex wikipedia , lookup

Synaptogenesis wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Long-term depression wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Synaptic gating wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Neurotransmitter wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Neuromuscular junction wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Chemical synapse wikipedia , lookup

NMDA receptor wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Transcript
J Neurophysiol 97: 817– 823, 2007.
First published November 8, 2006; doi:10.1152/jn.00980.2006.
Glycine Binding Sites of Presynaptic NMDA Receptors May Tonically
Regulate Glutamate Release in the Rat Visual Cortex
Yan-Hai Li and Tai-Zhen Han
Department of Physiology, School of Medicine, Xi’an Jiaotong University, Zhuque Dajie 205, Xi’an, Shannxi, People’s Republic of China
Submitted 13 September 2006; accepted in final form 4 November 2006
INTRODUCTION
Johnson and Ascher (1987) first demonstrated that, in the
vertebrate CNS, exogenously applied glycine potentiates the
response of N-methyl-D-aspartate receptors (NMDA-Rs) to
NMDA. Binding of glycine to the glycine binding site of
NMDA-Rs is a prerequirement for NMDA-R activation by
glutamate (Kleckner and Dingledine 1988). Thus glycine and
glutamate are co-agonists for NMDA-Rs.
It is generally assumed that the glycine binding site is
constantly saturated due to high concentrations of glycine in
cerebrospinal fluid (CSF) (Ferraro and Hare 1985). However,
glycine in the synaptic cleft is subject to a powerful uptake
system (i.e., glycine transporter 1 and 2, GlyT1 and GlyT2)
(Bergeron et al. 1998; Betz et al. 2006; Chen et al. 2003;
Supplisson and Bergman 1997). In addition, NMDA-R isoforms exhibit different affinities to glycine (Kew et al. 1998;
Kutsuwada et al. 1992). A surge of in vivo and in vitro studies
have suggested that saturation of the glycine binding varies
among different brain areas. Some studies have shown that the
application of exogenous glycine increases NMDA-R responses (Ahmadi et al. 2003; Baptista and Varanda 2005;
Berger et al. 1998; Czepita et al. 1996), whereas others have
found no modulatory effect (Fletcher et al. 1989; Obrenovitch
et al. 1997).
D-serine, which is a potent agonist of the glycine binding site
(Kleckner and Dingledine 1988; Martineau et al. 2006; Schell
et al. 1995) but is not taken up by glycine transporters (SupAddress for reprint requests and other correspondence: T.-Z. Han, Dept. of
Physiology, School of Medicine, Xi’an Jiaotong University, Zhuque Dajie 205,
Xi’an, Shaanxi 710061, PR China (E-mail: htzhen@@mail.xjtu.edu.cn).
www.jn.org
plisson and Bergman 1997), has been shown to potentiate
NMDA synaptic currents (Baptista and Varanda 2005; Mothet
et al. 2000). The ability of glycine or D-serine to potentiate
NMDA-R responses clearly indicates that the glycine binding
site of NMDA-Rs is not saturated.
To date, a number of studies have shown that the NMDA-Rs
are present on the presynaptic membrane in addition to the
postsynaptic area. Presynaptic NMDA-Rs have been found to
enhance neurotransmitter release in different regions of the
CNS including the spinal cord, cerebellum, entorhinal cortex,
and neocortex (Bardoni et al. 2004; Berretta and Jones 1996;
DeBiasi et al. 1996; Glitsch and Marty 1999; MacDermott et
al. 1999; Yang et al. 2006). They may also mediate synaptic
plasticity (i.e., LTP and LTD) (Casado et al. 2002; Duguid and
Sjostrom 2006; Sjostrom et al. 2003). Immunocytochemical
studies have demonstrated that presynaptic NR1 and NR2Bcontaining NMDA-Rs are present in rat visual cortex (Aoki et
al. 1994). Moreover, presynaptic NMDA-Rs enhance neurotransmitter release in the visual cortical layer V neurons as
proven by application of APV, a competitive antagonist of
these receptors (Sjostrom et al. 2003). However, whether the
glycine binding site of presynaptic NMDA-Rs may modulate
glutamatergic neurotransmission in CNS is still uncertain.
In this study, we examined whether the glycine binding site of
presynaptic NMDA-Rs regulates glutamate release in the rat
visual cortex. Our results show that the glycine binding site on the
postsynaptic NMDA-Rs is not saturated and that glycine binding
sites at presynaptic NMDA-Rs regulate glutamate release in the
layer II/III pyramidal neurons of the rat visual cortex.
METHODS
Slice preparation
Visual cortical slices were prepared from Sprague–Dawley rats
aged 13–15 days. All animals were housed in a standard environment
on a 12/12-h light/dark cycle with light on at 07:00, and they were
allowed ad libitum access to water and food. The use and care of
animals in this study follow the guidelines of the Xi’an Jiaotong
University Animal Research Advisory Committee. Rats were initially
anesthetized with ether and then immersed in ice-cold water, with the
nose exposed, for 3 min to reduce brain temperature. Immediately
after decapitation, the brain was placed in an ice-cold artificial
cerebrospinal fluid (ACSF) bubbled with 95% O2-5% CO2 (pH 7.4).
The ACSF consisted of (in mM) 124 NaCl, 5 KCl, 1.2 KH2PO4, 1.3
MgSO4, 2.4 CaCl2, 26 NaHCO3, and 10 glucose, and it had an
osmolality of 305–310 mosM/kg H2O. A block of tissue containing
the primary visual cortex was cut into 350-␮m-thick slices with a
vibratome (Campden Instruments, London, UK). Slices were transThe costs of publication of this article were defrayed in part by the payment
of page charges. The article must therefore be hereby marked “advertisement”
in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
0022-3077/07 $8.00 Copyright © 2007 The American Physiological Society
817
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
Li Y-H, Han T-Z. Glycine binding sites of presynaptic NMDA
receptors may tonically regulate glutamate release in the rat visual
cortex. J Neurophysiol 97: 817– 823, 2007. First published November
8, 2006; doi:10.1152/jn.00980.2006. In the CNS, activation of Nmethyl-D-aspartate receptor (NMDA-R) glycine binding sites is a
prerequisite for activation of postsynaptic NMDA-Rs by the excitatory neurotransmitter glutamate. Here we provide electrophysiological
evidence that the glycine binding sites of presynaptic NMDA-Rs
regulate glutamate release in layer II/III pyramidal neurons of the rat
visual cortex. Specifically, our results reveal that the frequency of
miniature excitatory postsynaptic currents is significantly reduced by
7-chloro-kynurenic acid (7-Cl KYNA), a NMDA-R glycine binding
site antagonist, and glycine or D-serine reverses this effect. Similar
results are obtained when the open-channel NMDA receptor blocker,
MK-801, is included in the recording pipette. Our data indicate that
the glycine binding site of postsynaptic NMDA-Rs is not saturated.
Moreover, they suggest that presynaptic NMDA-Rs are located in
layer II/III pyramidal neurons of the rat visual cortex and that the
glycine binding site of presynaptic NMDA-Rs tonically regulates
glutamate release.
818
Y.-H. LI AND T.-Z. HAN
ferred to an incubating chamber containing ACSF equilibrated with
carbogen (95% O2-5% CO2) and incubated for ⱖ1.5 h at room
temperature (20°C) prior to electrophysiological recording.
Recordings
Data analysis
Miniature excitatory postsynaptic currents (mEPSCs) were detected
and analyzed with MiniAnalysis software (Version 6.0, Synaptosoft)
using a threshold-crossing criterion. Although the threshold level
varied from neuron to neuron, it was always the same before and after
drug application (Bradaia and Trouslard 2002). Detection criteria also
included rise time ⬍3 ms (Sjostrom et al. 2003), which may eliminate
the purely NMDA-R-mediated mEPSCs because the average 10 –90%
rise time of purely NMDA-R-mediated mEPSC events is 8 ⫾ 2 ms
(O’Brien et al. 1997). This means that our data do not include the
silent synapses. Multiple peak events were discarded. Averaged
mEPSCs were obtained by aligning events on the rising phase. Events
occurring during 120 s were averaged and analyzed in every group. To
J Neurophysiol • VOL
RESULTS
Glycine or D-serine potentiates NMDA-R–mediated mEPSCs
Whole cell mEPSCs recordings were performed on 58 pyramidal neurons in layer II/III of the visual cortical slices. Spontaneous mEPSCs in the presence of TTX (0.5 ␮M), strychnine (1
␮M), and picrotoxin (100 ␮M) were abolished by application of
D-APV (50 ␮M) and CNQX (10 ␮M) in the Mg-free ACSF (data
not shown). This indicates that the mEPSCs were composed of
both NMDA and non-NMDA glutamatergic components (Berger
et al. 1998; O’Brien et al. 1997). Application of exogenous
glycine (100 ␮M) enhanced the NMDA-R-mediated mEPSCs,
and this effect was completely blocked by D-APV (50 ␮M; Fig. 1,
A and B). This enhancement of NMDA-R-mediated mEPSCs or
increased NMDA charge transfer (fC), is shown in Fig. 1C.
Charge transfers under control conditions and in response to
glycine treatment were 622 ⫾ 111 and 795 ⫾ 104 fC, respectively
(28% increase, P ⬍ 0.007, n ⫽ 8). These results indicate that the
glycine binding site on synaptically active NMDA-Rs is not
saturated. However, application of glycine (100 ␮M) did not alter
the frequency or the peak amplitude of the mEPSCs (Fig. 1, D and
E). Application of D-APV (50 ␮M) decreased the frequency
(control, 2.82 ⫾ 0.5 Hz vs. treated, 1.50 ⫾ 0.5 Hz, P ⬍ 0.001, n ⫽
8) but not the peak amplitude of the mEPSCs (control, 46.9 ⫾ 4.4
pA vs. treated, 46.0 ⫾ 4.6 pA, P ⬎ 0.18, n ⫽ 8). This result
suggests that glycine binding site on presynaptic NMDA-Rs is
saturated and the NMDA-Rs may tonically regulate glutamate
release.
D-serine, another agonist at the glycine-site of NMDA-Rs, is
not transported by glycine transporters (Supplisson and Bergman 1997). Therefore we examined the effect of D-serine on
NMDA-mediated mEPSCs to verify the results obtained with
glycine. Application of exogenous D-serine (100 ␮M) also
increased the NMDA-R-mediated component of mEPSCs,
which was abolished by D-APV (50 ␮M; Fig. 2, A and B).
NMDA-mediated charge transfer increased from 631 ⫾ 83 fC
in the control to 960 ⫾ 186 fC in D-serine-treated neurons
(52% increase, P ⬍ 0.005, n ⫽ 9, Fig. 2C). Thus D-serine had
a greater effect on the glycine binding site of NMDA-Rs than
glycine. However, D-serine did not alter the frequency or the
peak amplitude of the mEPSCs (Fig. 2, D and E). Application
of D-APV reduced the frequency (control, 2.27 ⫾ 0.6 Hz vs.
treated, 0.98 ⫾ 0.3, P ⬍ 0.001, n ⫽ 9) but not the peak
amplitude (control, 40.3 ⫾ 3.6 pA vs. treated, 39.4 ⫾ 3.5 pA,
P ⬎ 0.21, n ⫽ 9) of the mEPSCs. These results verified that
D-serine has the same effect as glycine.
Glycine or D-serine reverses the inhibitory effects of 7-Cl
KYNA on the frequency of NMDA-R–mediated mEPSCs
To determine whether the glycine binding site of the
NMDA-Rs may be involved in the modulation of glutamate
97 • JANUARY 2007 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
For recording, slices were individually transferred to a recording
chamber where they were perfused (2.5–3 ml/min) with oxygenated
Mg-free ACSF at 31 ⫾ 0.5°C. The temperature of the recording
chamber was continuously monitored and controlled by a custommade temperature controller. The slices were placed on an upright
infra-red video microscope with differential interference contrast
(DIC) optics (OLYMPUS BX51WI), which was mounted on a
Gibraltar X–Y table. Slices were observed through a ⫻40 waterimmersion objective using an infrared-sensitive camera (DAGE-MTI,
IR-1000). Layer II/III pyramidal neurons of the visual cortex were
visually selected (Mason and Larkman 1990). Patch-clamp recordings
were performed in the whole cell configuration. Unpolished and
uncoated patch pipettes (1.5 mm/1.1 mm; Sutter Instruments, Novato,
CA) with a resistance of 4 – 6 M⍀ were pulled using a horizontal
puller (model P-97, Sutter Instruments, Novato, CA). The pipette
solution contained (in mM) 130 cesium methanesulfonate, 5 NaCl, 5
QX-314, 2 MgCl2, 0.1 CaCl2, 1 EGTA, 10 HEPES, 2 Na2ATP, and
0.25 Na3GTP (pH 7.3–7.4, adjusted with CsOH, 280 –290 mosM/kg
H2O). To block postsynaptic NMDA-Rs, the following pipette solution was used (in mM): 124 cesium methanesulfonate, 5 QX-314, 2
MgCl2, 10 BAPTA, 1 (⫹)-MK-801, 10 HEPES, 2 Na2ATP, and 0.25
Na3GTP (pH 7.3–7.4, adjusted with CsOH, 280 –290 mosM/kg H2O).
To facilitate the blockade of the open channel blocker MK-801,
neurons were depolarized to ⫺10 mV for 10 s at intervals during a
10-min period after breakthrough to whole cell access, and this
interval might be enough for MK-801 to block all the channels
according to Yang’s work (Yang et al. 2006). Currents were recorded
at a holding potential of ⫺60 mV using an Axopatch 700B amplifier
(Axon Instruments, USA) and digitized using a data-acquisition board
(Digidata 1322A) operated by pCLAMP 9.2 software (Axon Instruments). They were filtered at 2 kHz and digitized at 10 kHz. Compensation for the series resistance was not employed. Statistical
comparisons were performed only when series resistance was ⬍20
M⍀ and did not change by ⬎10%.
Glycine, D-serine, strychnine, picrotoxin, 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX), D-2-amino-5-phosphonovalerate (D-APV),
(⫹)-MK-801, QX-314, BAPTA, and 7-chloro-kynurenic acid (7-Cl
KYNA) were purchased from Sigma. Tetrodotoxin (TTX) was obtained from Hebei Province Marine Science and Technolog, China.
TTX (0.5 ␮M), strychnine (1 ␮M) and picrotoxin (100 ␮M) were
added to Mg-free ACSF to suppress the sodium currents, glycinergic
and GABAergic transmission, respectively. Mg-free ACSF was obtained by omitting MgSO4 (1.3 mM) from the ACSF without compensation for the loss of osmolarity or for the amount of divalent ions.
All drugs were applied at known concentrations by changing the
perfusion line.
quantify the NMDA component of mEPSCs, we integrated the average mEPSCs waveforms occurring over a 90-ms time period, which
began 10 ms after the start of the mEPSCs (Berger et al. 1998). This
integral was used as an index of NMDA-R-mediated charge transfer.
Comparison of mEPSC distributions was performed using the Kolmogorov-Smirnov test, and values of P ⬍ 0.01 were accepted as
significant (Berretta and Jones 1996). Group means were compared
using a paired Student’s t-test, and P ⬍ 0.05 was considered significant. All data are expressed as means ⫾ SE.
GLYCINE TONICALLY REGULATES GLUTAMATE RELEASE
819
release, we employed 7-Cl KYNA, an antagonist of the glycine
binding site of the NMDA-Rs. As shown in Fig. 3D, inclusion
of 7-Cl KYNA (20 ␮M) in the extracellular solution blocked
the postsynaptic NMDA component and reduced the frequency
of the mEPSCs (control, 1.9 ⫾ 0.7 Hz vs. treated, 1.1 ⫾ 0.4
Hz, P ⬍ 0.001, n ⫽ 10). Combined treatment with glycine (100
␮M) and 7-Cl KYNA reversed this effect (control, 1.9 ⫾ 0.7
Hz vs. treated, 1.8 ⫾ 0.6 Hz, P ⬎ 0.08, n ⫽ 10) and recovered
the postsynaptic NMDA component (Fig. 3, B and D). As
shown in Fig. 3C, the peak amplitude of the mEPSCs was not
influenced by 7-Cl KYNA or glycine (control, 41.0 ⫾ 3.3 pA;
7-Cl KYNA, 39.6 ⫾ 4.0 pA; 7-Cl KYNA ⫹ glycine, 40.4 ⫾
3.9 pA, n ⫽ 10).
As shown in Fig. 4, D-serine (100 ␮M) also reversed the
inhibitory effects of 7-Cl KYNA (20 ␮M) on the frequency of
NMDA-mediated mEPSCs (control, 1.69 ⫾ 0.4 Hz; 7-Cl
KYNA, 1.1 ⫾ 0.2 Hz; 7-Cl KYNA ⫹ D-serine, 1.65 ⫾ 0.3 Hz,
P ⬍ 0.001, n ⫽ 12). The peak amplitude of the mEPSCs was
unaffected by treatment either with 7-Cl KYNA only or along
with D-serine (control, 40 ⫾ 3.8 pA; 7-Cl KYNA, 38.6 ⫾ 5.1
pA; 7-Cl KYNA ⫹ D-serine, 39.0 ⫾ 4.4 pA; n ⫽ 12; Fig. 4C).
These results suggest that the glycine binding site is involved
in the modulation of the glutamate release.
Glycine or D-serine reverses the inhibitory effects of 7-Cl
KYNA in the presence of MK-801
It is possible that 7-Cl KYNA reduced the frequency of the
mEPSCs by acting on the glycine binding site of presynaptic
NMDA-Rs or postsynaptic NMDA-Rs. In the latter case,
J Neurophysiol • VOL
presynaptic neurotransmitter release could be affected through
retrograde messengers. To test these possibilities, we included
the open-channel NMDA-R blocker, dizocilpine maleate (MK801; 1 mM), in the recording pipette as described by Berretta
and Jones (1996). Blockade of the postsynaptic NMDA-Rs
with MK-801 completely abolished the NMDA component of
mEPSCs (Fig. 5). Inclusion of 7-Cl KYNA (20 ␮M) under
these conditions reduced the frequency of the mEPSCs. Moreover, this effect could be reversed by the addition of glycine
(100 ␮M; control, 2.6 ⫾ 0.5 Hz; 7-Cl KYNA, 1.7 ⫾ 0.3 Hz;
7-Cl KYNA ⫹ glycine, 2.6 ⫾ 0.6 Hz; P ⬍ 0.001, n ⫽ 10; Fig.
5C). As before, the peak amplitude of the mEPSCs was not
affected by treatment with 7-Cl KYNA and glycine. (control,
34.3 ⫾ 1.5 pA; 7-Cl KYNA, 34.4 ⫾ 1.2 pA; 7-Cl KYNA ⫹
glycine, 34.5 ⫾ 1.7 pA; n ⫽ 10).
As shown in Fig. 6C, D-serine (100 ␮M) also reversed the
effects of the 7-Cl KYNA on mEPSC frequency in the presence of MK-801 (control, 2.8 ⫾ 0.7 Hz; 7-Cl KYNA, 1.6 ⫾ 0.5
Hz; 7-Cl KYNA ⫹ D-serine, 2.8 ⫾ 0.6 Hz; P ⬍ 0.001; n ⫽ 9).
The peak amplitude of the mEPSCs was not affected by these
treatments (control, 32.8 ⫾ 3.9 pA; 7-Cl KYNA, 32.4 ⫾ 4.1
pA; 7-Cl KYNA ⫹ D-serine, 33.1 ⫾ 4.2 pA; n ⫽ 9; Fig. 6B).
These results suggest that presynaptic, but not postsynaptic,
NMDA-R glycine binding sites tonically regulate glutamate
release in the visual cortex.
Additional experiments showed that there were no effects of
glycine, D-serine or 7-Cl KYNA on the frequency and amplitude of mEPSCs in the presence of D-APV, which blocked both
pre- and postsynaptic NMDARs (data showing in Supplemental materials).
97 • JANUARY 2007 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
FIG. 1. Glycine enhances N-methyl-D-aspartate receptor
(NMDA-R)-mediated miniature excitatory postsynaptic
currents (mEPSCs). A: traces of mEPSCs recorded from the
same neuron (holding potential ⫽ ⫺60 mV) under control
conditions (top), after the addition of glycine (middle), and
after addition of D-APV (bottom). The NMDA component
of mEPSCs was increased by glycine and decreased by
D-APV. B: average mEPSCs from the same neuron shown
in A. C: NMDA charge transfer occurring between 10 and
100 ms after mEPSC onset was increased in the presence of
100 ␮M glycine (P ⬍ 0.007, n ⫽ 8). D: pooled data from
8 neurons shows the peak amplitudes of the mEPSCs were
not altered by treatment with glycine and D-APV. E: glycine (100 ␮M) did not alter the frequency of the mEPSCs.
However, application of D-APV (50 ␮M) decreased the
frequency (P ⬍ 0.001, n ⫽ 8).
820
Y.-H. LI AND T.-Z. HAN
DISCUSSION
In the present study, we have used a whole cell recording
technique to examine the effects of glycine and D-serine on the
spontaneous synaptic activities in layer II/III pyramidal neurons of the rat visual cortex. All data were obtained in the
presence of TTX (0.5 ␮M), strychnine (1 ␮M), and picrotoxin
(100 ␮M) in the Mg-free ACSF and at a holding potential of
⫺60 mV. Under these conditions, the spontaneous spikes as
well as glycinergic and GABAergic synaptic activities could be
excluded, and the NMDA-R- mediated mEPSCs could be
FIG. 3. Glycine reverses the inhibitory effect of
7-chloro-kynurenic acid (7-Cl KYNA) on NMDA-R-mediated mEPSCs. A: traces of mEPSCs recorded from the
same neuron (holding potential ⫽ ⫺60 mV) under control
conditions (top) and in the presence of only 20 ␮M 7-Cl
KYNA (middle) and along with 100 ␮M glycine (bottom).
B: average mEPSCs from the same neuron shown in A. C:
pooled data from 10 neurons shows the peak amplitudes of
the mEPSCs were not altered by treatment with 7-Cl
KYNA and glycine. D: frequency of mEPSCs was reduced
by 7-Cl KYNA (P ⬍ 0.001, n ⫽ 10), and this effect was
reversed by addition of glycine (P ⬎ 0.08, n ⫽ 10).
J Neurophysiol • VOL
97 • JANUARY 2007 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
FIG. 2. D-serine enhances the NMDA-R-mediated
mEPSCs. A: traces of mEPSCs recorded from the same
neuron (holding potential ⫽ ⫺60 mV) under control
conditions (top), after the addition of D-serine (middle),
and after addition of D-APV (bottom). The NMDA
component of mEPSCs was increased by D-serine and
decreased by D-APV. B: average mEPSCs from the
same neuron shown in A. C: NMDA charge transfer
occurring between 10 and 100 ms after mEPSC onset
was increased in the presence of D-serine (P ⬍ 0.005,
n ⫽ 9). D: pooled data from nine neurons shows the
peak amplitudes of the mEPSCs were not altered by
treatment with D-serine and D-APV. E: D-serine (100
␮M) did not alter the frequency of the mEPSCs. However, application of D-APV (50 ␮M) decreased the
frequency (P ⬍ 0.001, n ⫽ 9).
GLYCINE TONICALLY REGULATES GLUTAMATE RELEASE
821
facilitated. The finding that mEPSCs were abolished by the
presence of D-APV (50 ␮M) and CNQX (10 ␮M) in extracellular fluid indicates that the mEPSCs were composed of both
NMDA and non-NMDA glutamatergic components. On the
other hand, mEPSCs were made up of only non-NMDA
glutamatergic components when MK-801 was included in the
recording pipette.
As described in the preceding text, application of exogenous
glycine potentiated NMDA-R-mediated mEPSCs. Glycine
may achieve this effect through at least three mechanisms.
First, glycine might act through presynaptic, strychnine-sensitive glycine receptors to facilitate the release of glutamate, an
action that has been demonstrated in the brain stem (Turecek
and Trussell 2001). In this study, the contribution of such as
mechanism is probably limited because mEPSCs were recorded in the presence of strychnine (1 ␮M) and picrotoxin
(100 ␮M). Second, glycine might act on the NMDA-Rs containing NR3A or NR3B subunits (Chatterton et al. 2002),
which are activated by glycine in the absence of glutamate. The
finding that D-serine, which blocks NMDA-Rs containing
NR3A or NR3B subunits, potentiates NMDA-R–mediated
mEPSCs argues against this possibility. Third, glycine might
act postsynaptically by activating the glycine binding site of
NMDA-Rs. To test this hypothesis, we employed D-serine,
another NMDA-R glycine binding site agonist. The ability of
D-serine to increase the NMDA component provides evidence
that glycine potentiates NMDA-R–mediated mEPSCs via the
NMDA-R glycine binding site. In addition, the non-NMDA
FIG. 5. Glycine reverses the inhibitory effects of
7-Cl KYNA on mEPSC frequency in the presence of
MK-801. A: NMDA-R open channel blocker, MK-801
(1 mM), was applied via the recording pipette, and
mEPSCs were recorded under control conditions (top)
and in the presence of only 20 ␮M 7-Cl KYNA
(middle) and along with 100 ␮M glycine (bottom).
The mEPSC tracings from the same neuron (holding
potential ⫽ ⫺60 mV) are shown. Blockade of
postsynaptic NMDA-Rs via MK-801 completely abolished NMDA-mediated mEPSCs. B: average mEPSCs
shown in A. C: pooled data from 10 neurons shows the
peak amplitudes of the mEPSCs were not significantly
changed under any of these conditions. D: in the
presence of MK-801, the frequency of mEPSCs was
reduced by 7-Cl KYNA, and this effect was reversed
by ddition of glycine (P ⬍ 0.001, n ⫽ 10).
J Neurophysiol • VOL
97 • JANUARY 2007 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
FIG. 4. D-serine reverses the inhibitory effect of 7-Cl
KYNA on NMDA-R-mediated mEPSCs. A: traces of
mEPSCs recorded from the same neuron (holding potential ⫽ ⫺60 mV) under control conditions (top) and in the
presence of 20 ␮M 7-Cl KYNA (middle) and along with
100 ␮M D-serine (bottom). B: average mEPSCs from the
same neuron shown in A. C: pooled data from 12 neurons
shows the peak amplitudes of the mEPSCs were not
altered by treatment with 7-Cl KYNA and D-serine. D:
frequency of mEPSCs was reduced by 7-Cl KYNA, and
this effect was reversed by addition of D-serine (P ⬍
0.001, n ⫽ 12).
822
Y.-H. LI AND T.-Z. HAN
component was not influenced by glycine or D-serine, and the
application of APV or 7-Cl KYNA abolished the NMDA
component of the mEPSCs. Taken together, these results suggest that the glycine binding site on the postsynaptic
NMDA-Rs is not saturated as has been shown in cat visual
cortex (Czepita et al. 1996).
The levels of synaptic glycine are known to be tightly
controlled by glycine transporters. There are two types of
glycine transporters, GlyT1 and GlyT2. GlyT1, which is localized to glia and neurons, is closely associated with the
NMDA-Rs (Cubelos et al. 2005); whereas, GlyT2 is colocalized with strychnine-sensitive glycine receptors (Gomeza et al.
2003). As has been described in the hippocampus (Watanabe et
al. 1992), D-serine was capable of potentiating NMDA synaptic
currents. Here, application of exogenous D-serine increased the
NMDA component by ⬃52%, and glycine increased the
NMDA component of the mEPSCs by⬃ 28%. The greater
effect of D-serine, which is not taken up by glycine transporters
(Supplisson and Bergman 1997), suggests that levels of synaptic glycine are relatively low and are tightly controlled by
glycine transporters.
Our results also show that APV or 7-Cl KYNA reduce the
frequency but not the peak amplitude of mEPSCs, regardless of
the presence of MK-801 in the recording pipette. These findings support the existence of presynaptic NMDA-Rs and are
consistent with studies showing that presynaptic NMDA-Rs
tonically facilitate glutamate release as a result of increased
calcium influx into the terminals (Berretta and Jones 1996;
Sjostrom et al. 2003; Yang et al. 2006).
It was reported that mEPSCs events rely on presynaptic
Ca2⫹ transient either in hippocamcal area CA1 or in layer II
neurons of the rat barrel cortex (Rusakov 2006; Simkus and
Stricker 2002). Presynaptic Ca2⫹ transient could produce Ca2⫹
release from stores, named Ca2⫹-induced Ca2⫹ release
J Neurophysiol • VOL
(CICR), and then the postsynaptic EPSCs. The amplitude of
CICR from internal Ca2⫹ stores depends on the amount of
Ca2⫹ influx and probably on the cytosolic Ca2⫹ concentration
(Rusakov 2006). Activation of presynaptic NMDA-Rs could
induce the Ca2⫹ influx and then improve the CICR and finally
facilitate the neurotransmitter release.
The difference in saturation of the glycine site between preand postsynaptic NMDA-Rs seems to have different affinities
for glycine (Kew et al. 1998; Kutsuwada et al. 1992). There is
strong evidence that at P14 the postsynaptic NMDA receptors
are mostly NR1-NR2A, whereas the presynaptic ones are
mostly NR1-NR2B (Sjostrom et al. 2003; Woodhall et al.
2001; Yang et al. 2006). The EC50s of glycine and of D-serine
for NR1-NR2A receptors are known to be higher than those for
NR1-NR2B. Thus a concentration of glycine will be saturating
for the (presynaptic) NR2B-containing receptors and not saturating for the (postsynaptic) NR2A-containing receptors. Also
consistent with our findings, presynaptic NR1- and NR2Bcontaining NMDA-Rs are known to exist (Aoki et al. 1994)
and enhance neurotransmitter release in the visual cortex (Sjostrom et al. 2003).
The ability of glycine or D-serine to reverse the effects of
7-Cl KYNA when MK-801 was present in the recording
pipette argues against the possibility that glycine acts on
postsynaptic receptors to facilitate glutamate release via retrograde messengers. Additionally, the finding that D-serine is
able to reverse the effects of 7-Cl KYNA, combined with the
fact that strychnine was present in all experiments, makes it
unlikely that presynaptic, strychnine-sensitive glycine receptors facilitated glutamate release (Turecek and Trussell 2001).
In summary, our data show that glycine can act on glycine
binding sites at presynaptic and postsynaptic NMDA-Rs to
enhance the NMDA-R function, which could be indicative of a
general role for the glycine binding site of presynaptic
97 • JANUARY 2007 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
FIG. 6. D-serine reverses the inhibitory effects of
7-Cl KYNA on mEPSCs frequency in the presence of
MK-801. A: NMDA-R open channel blocker, MK-801
(1 mM), was applied via the recording pipette, and
mEPSCs were recorded under control conditions (top)
and in the presence of only 20 ␮M 7-Cl KYNA
(middle) and along with 100 ␮M glycine (bottom).
The mEPSC tracings from the same neuron (holding
potential ⫽ ⫺60 mV) are shown. Blockade of
postsynaptic NMDA-Rs via MK-801 completely abolished NMDA-mediated mEPSCs. B: average mEPSCs
shown in A. C: pooled data from 9 neurons shows the
peak amplitudes of the mEPSCs were not significantly
changed under any of these conditions. D: in the
presence of MK-801, the frequency of mEPSCs was
reduced by 7-Cl KYNA, and this effect was reversed
by addition of D-serine (P ⬍ 0.001, n ⫽ 9).
GLYCINE TONICALLY REGULATES GLUTAMATE RELEASE
NMDA-Rs in regulating glutamate release in the CNS. Moreover, these findings may be clinically relevant to schizophrenia, where enhancing NMDA-R function is considered to be a
promising strategy for treatment of the disease (Coyle et al.
2002; Duncan et al. 1999).
GRANTS
This work was supported by the National Natural Science Fountation of
China Grant 30170310 and the key program of Xi’an Jiaotong University.
REFERENCES
J Neurophysiol • VOL
Ferraro TN, Hare TA. Free and conjugated amino acids in human CSF:
influence of age and sex. Brain Res 338: 53– 60, 1985.
Fletcher EJ, Millar JD, Zeman S, Lodge D. Non-competitive antagonism of
N-methyl-D-aspartate by displacement of an endogenous glycine-like substance. Eur J Neurosci 1: 196 –203, 1989.
Glitsch M, Marty A. Presynaptic effects of NMDA in cerebellar Purkinje
cells and interneurons. J Neurosci 19: 511–519, 1999.
Gomeza J, Ohno K, Betz H. Glycine transporter isoforms in the mammalian
central nervous system: structures, functions and therapeutic promises. Curr
Opin Drug Discov Dev 6: 675– 682, 2003.
Johnson JW, Ascher P. Glycine potentiates the NMDA response in cultured
mouse brain neurons. Nature 325: 529 –531, 1987.
Kew JN, Richards JG, Mutel V, Kemp JA. Developmental changes in
NMDA receptor glycine affinity and ifenprodil sensitivity reveal three
distinct populations of NMDA receptors in individual rat cortical neurons.
J Neurosci 18: 1935–1943, 1998.
Kleckner NW, Dingledine R. Requirement for glycine in activation of
NMDA-receptors expressed in Xenopus oocytes. Science 241: 835– 837,
1988.
Kutsuwada T, Kashiwabuchi N, Mori H, Sakimura K, Kushiya E, Araki
K, Meguro H, Masaki H, Kumanishi T, Arakawa M, Mishina M.
Molecular diversity of the NMDA receptor channel. Nature 358: 36 – 41,
1992.
MacDermott AB, Role LW, Siegelbaum SA. Presynaptic ionotropic receptors and the control of transmitter release. Annu Rev Neurosci 22: 443– 485,
1999.
Martineau M, Baux G, and Mothet JP. d-Serine signalling in the brain:
friend and foe. Trends Neurosci 29: 481– 491, 2006.
Mason A, Larkman A. Correlations between morphology and electrophysiology of pyramidal neurons in slices of rat visual cortex. II. Electrophysiology. J Neurosci 10: 1415–1428, 1990.
Mothet JP, Parent AT, Wolosker H, Brady RO Jr, Linden DJ, Ferris CD,
Rogawski MA, Snyder SH. D-serine is an endogenous ligand for the
glycine site of the N-methyl-D-aspartate receptor. Proc Natl Acad Sci USA
97: 4926 – 4931, 2000.
O’Brien JA, Isaacson JS, Berger AJ. NMDA and non-NMDA receptors are
co-localized at excitatory synapses of rat hypoglossal motoneurons. Neurosci Lett 227: 5– 8, 1997.
Obrenovitch TP, Hardy AM, Urenjak J. High extracellular glycine does not
potentiate N-methyl-D-aspartate-evoked depolarization in vivo. Brain Res
746: 190 –194, 1997.
Rusakov DA. Ca2⫹-dependent mechanisms of presynaptic control at central
synapses. Neuroscientist 12: 317–326, 2006.
Schell MJ, Molliver ME, Snyder SH. D-serine, an endogenous synaptic
modulator: localization to astrocytes and glutamate-stimulated release. Proc
Natl Acad Sci USA 92: 3948 –3952, 1995.
Simkus CR, Stricker C. The contribution of intracellular calcium stores to
mEPSCs recorded in layer II neurons of rat barrel cortex. J Physiol 545:
521–535, 2002.
Sjostrom PJ, Turrigiano GG, Nelson SB. Neocortical LTD via coincident
activation of presynaptic NMDA and cannabinoid receptors. Neuron 39:
641– 654, 2003.
Supplisson S, Bergman C. Control of NMDA receptor activation by a glycine
transporter co-expressed in Xenopus oocytes. J Neurosci 17: 4580 – 4590,
1997.
Turecek R, Trussell LO. Presynaptic glycine receptors enhance transmitter
release at a mammalian central synapse. Nature 411: 587–590, 2001.
Watanabe Y, Saito H, Abe K. Effects of glycine and structurally related
amino acids on generation of long-term potentiation in rat hippocampal
slices. Eur J Pharmacol 223: 179 –184, 1992.
Woodhall G, Evans DI, Cunningham MO, Jones RS. NR2B-containing
NMDA autoreceptors at synapses on entorhinal cortical neurons. J Neurophysiol 86: 1644 –1651, 2001.
Yang J, Woodhall GL, Jones RS. Tonic facilitation of glutamate release by
presynaptic NR2B-containing NMDA receptors is increased in the entorhinal cortex of chronically epileptic rats. J Neurosci 26: 406 – 410, 2006.
97 • JANUARY 2007 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.2 on August 3, 2017
Ahmadi S, Muth-Selbach U, Lauterbach A, Lipfert P, Neuhuber WL,
Zeilhofer HU. Facilitation of spinal NMDA receptor currents by spillover
of synaptically released glycine. Science 300: 2094 –2097, 2003.
Aoki C, Venkatesan C, Go CG, Mong JA, Dawson TM. Cellular and
subcellular localization of NMDA-R1 subunit immunoreactivity in the
visual cortex of adult and neonatal rats. J Neurosci 14: 5202–5222, 1994.
Baptista V, Varanda WA. Glycine binding site of the synaptic NMDA
receptor in subpostremal NTS neurons. J Neurophysiol 94: 147–152, 2005.
Bardoni R, Torsney C, Tong CK, Prandini M, MacDermott AB. Presynaptic NMDA receptors modulate glutamate release from primary sensory
neurons in rat spinal cord dorsal horn. J Neurosci 24: 2774 –2781, 2004.
Berger AJ, Dieudonne S, Ascher P. Glycine uptake governs glycine site
occupancy at NMDA receptors of excitatory synapses. J Neurophysiol 80:
3336 –3340, 1998.
Bergeron R, Meyer TM, Coyle JT, Greene RW. Modulation of N-methylD-aspartate receptor function by glycine transport. Proc Natl Acad Sci USA
95: 15730 –15734, 1998.
Berretta N, Jones RS. Tonic facilitation of glutamate release by presynaptic
N-methyl-D-aspartate autoreceptors in the entorhinal cortex. Neuroscience
75: 339 –344, 1996.
Betz H, Gomeza J, Armsen W, Scholze P, Eulenburg V. Glycine transporters: essential regulators of synaptic transmission. Biochem Soc Trans 34:
55–58, 2006.
Bradaia A, Trouslard J. Nicotinic receptors regulate the release of glycine
onto lamina X neurons of the rat spinal cord. Neuropharmacology 43:
1044 –1054, 2002.
Casado M, Isope P, Ascher P. Involvement of presynaptic N-methyl-Daspartate receptors in cerebellar long-term depression. Neuron 33: 123–130,
2002.
Chatterton JE, Awobuluyi M, Premkumar LS, Takahashi H, Talantova
M, Shin Y, Cui J, Tu S, Sevarino KA, Nakanishi N, Tong G, Lipton SA,
Zhang D. Excitatory glycine receptors containing the NR3 family of
NMDA receptor subunits. Nature 415: 793–798, 2002.
Chen L, Muhlhauser M, Yang CR. Glycine tranporter-1 blockade potentiates
NMDA-mediated responses in rat prefrontal cortical neurons in vitro and in
vivo. J Neurophysiol 89: 691–703, 2003.
Coyle JT, Tsai G, Goff DC. Ionotropic glutamate receptors as therapeutic
targets in schizophrenia. Curr Drug Targets CNS Neurol Disord 1: 183–189,
2002.
Cubelos B, Gimenez C, Zafra F. Localization of the GLYT1 glycine
transporter at glutamatergic synapses in the rat brain. Cereb Cortex 15:
448 – 459, 2005.
Czepita D, Daw NW, Reid SN. Glycine at the NMDA receptor in cat visual
cortex: saturation and changes with age. J Neurophysiol 75: 311–317, 1996.
DeBiasi S, Minelli A, Melone M, Conti F. Presynaptic NMDA receptors in
the neocortex are both auto- and heteroreceptors. Neuroreport 7: 2773–
2776, 1996.
Duguid I, Sjostrom PJ. Novel presynaptic mechanisms for coincidence
detection in synaptic plasticity. Curr Opin Neurobiol 16: 312–322, 2006.
Duncan GE, Sheitman BB, Lieberman JA. An integrated view of pathophysiological models of schizophrenia. Brain Res Brain Res Rev 29:
250 –264, 1999.
823