Sei sulla pagina 1di 14

Cerebral Cortex, 2019;00: 1–14

doi: 10.1093/cercor/bhz174
Advance Access Publication Date:
Original Article

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


ORIGINAL ARTICLE

Synapsin I Synchronizes GABA Release in Distinct


Interneuron Subpopulations
N. Forte1,2,† , F. Binda1,2 , A. Contestabile3 , F. Benfenati1,2 and P. Baldelli2,4
1 Centerfor Synaptic Neuroscience and Technology, Istituto Italiano di Tecnologia, Largo Rosanna Benzi 10,
16132 Genova, Italy, 2 IRCSS, Ospedale Policlinico San Martino, Largo Rosanna Benzi 10, 16132 Genova, Italy,
3 Department of Neuroscience and Brain Technologies, Istituto Italiano di Tecnologia, Via Morego 30, 16163

Genova, Italy and 4 Department of Experimental Medicine, University of Genova, Viale Benedetto XV, 3, 16132
Genova, Italy
Address correspondence to Fabio Benfenati, Center for Synaptic Neuroscience and Technology, Istituto Italiano di Tecnologia, Largo Rosanna Benzi 10,
16132 Genova, Italy. Email: fabio.benfenati@iit.it
† Present address: Endocannabinoid Research Group, Institute of Biomolecular Chemistry, CNR, Viale Campi Flegrei, 34, 80078 Pozzuoli (NA), Italy

Benfenati F. and Baldelli P. contributed equally to this study

Abstract
Neurotransmitters can be released either synchronously or asynchronously with respect to action potential timing.
Synapsins (Syns) are a family of synaptic vesicle (SV) phosphoproteins that assist gamma-aminobutyric acid (GABA) release
and allow a physiological excitation/inhibition balance. Consistently, deletion of either or both Syn1 and Syn2 genes is
epileptogenic. In this work, we have characterized the effect of SynI knockout (KO) in the regulation of GABA release
dynamics. Using patch-clamp recordings in hippocampal slices, we demonstrate that the lack of SynI impairs synchronous
GABA release via a reduction of the readily releasable SVs and, in parallel, increases asynchronous GABA release. The
effects of SynI deletion on synchronous GABA release were occluded by ω-AgatoxinIVA, indicating the involvement of
P/Q-type Ca2+ channel-expressing neurons. Using in situ hybridization, we show that SynI is more expressed in
parvalbumin (PV) interneurons, characterized by synchronous release, than in cholecystokinin or SOM interneurons,
characterized by a more asynchronous release. Optogenetic activation of PV and SOM interneurons revealed a specific
reduction of synchronous release in PV/SynIKO interneurons associated with an increased asynchronous release in
SOM/SynIKO interneurons. The results demonstrate that SynI is differentially expressed in interneuron subpopulations,
where it boosts synchronous and limits asynchronous GABA release.

Key words: GABA interneurons, hippocampus, parvalbumin, somatostatin, Synapsin I

Introduction release represents a “delayed” release phase generated by


Neurotransmitter release is a multistep and highly regulated residual Ca2+ that slowly recruits groups of synaptic vesicles
process based on the participation of hundreds of tightly (SVs) far from the Ca2+ sources (Kaeser and Regehr 2014). These
regulated proteins. Alterations of this sophisticated machinery two release modalities characterize distinct subpopulations
lead to defects in synaptic transmission and consequently to of interneurons: parvalbumin (PV)-positive interneurons are
the development of neurological diseases such as epilepsy markedly synchronous in releasing γ -aminobutyric acid (GABA),
or intellectual disability. Two distinct modalities of synaptic while cholecystokinin (CCK)- and somatostatin (SOM)-positive
communication exist, differently coupled in time to the action interneurons predominantly release GABA asynchronously
potential (AP) invasion of the presynaptic terminal. The syn- (Hefft and Jonas 2005; Savanthrapadian et al. 2014; Liguz-Lecznar
chronous release is tightly coupled to the Ca2+ rise in the close et al. 2016; Tremblay et al. 2016). Both phasic synaptic inhibition,
vicinity of Ca2+ channels, at the active zone, while asynchronous derived from synchronous GABA release, and tonic shunting

© The Author(s) 2019. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com
2 Cerebral Cortex, 2019, Vol. 00, No. 00

inhibition, largely contributed by asynchronous GABA release Materials and Methods


(Medrihan et al. 2015), cooperate for the spatial and temporal
Experimental Animals
control of network activity. Alterations in the synchronous/asyn-
chronous release ratio (s/a-R) strongly impact on network All experiments were performed on 3–7 weeks old male

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


excitability and can trigger epileptogenesis (Jiang et al. 2012; C57BL/6J, SynIKO, PV-cre, SOM-cre, PV-cre/SynIKO, or SOM-
Medrihan et al. 2015). cre/SynIKO. C57BL/6J were obtained from Charles River (Calco,
The synapsins (Syns) compose a family of presynaptic phos- Italy), SynIKO mice were generated by homologous recombi-
phoproteins encoded by three distinct genes (Syn1, Syn2, and nation and extensively backcrossed on a C57BL/6J background,
Syn3). By interacting with SVs and the actin cytoskeleton, Syns PV-cre and SOM-cre mice (Taniguchi et al. 2011) were obtained
are involved in both predocking and postdocking stages of neu- from Jackson Laboratories (Bar Harbor, ME) and backcrossed
rotransmitter release, regulating SV trafficking and short-term onto the C57BL/6J background. Since Syn1 is X-linked, we
plasticity (Cesca et al. 2010). Mutations in the human SYN1 and crossed male PV-cre or SOM-cre homozygous mice with either
SYN2 genes were reported to associate with epilepsy, autism C57BL/6J or SynIKO females to obtain single-generation PV-
spectrum disorder, or intellectual disability (Garcia 2004; Fassio cre/SynIKO or SOM-cre/SynIKO hemizygous male mice. For
et al. 2011a; Giannandrea et al. 2013; Lignani et al. 2013; Corradi in situ hybridization experiments, male PV-Tomato Red and
et al. 2014; Nguyen et al. 2015; Guarnieri et al. 2017; Peron SOM Tomato Red were obtained from Jackson Laboratories,
et al. 2018). Similarly, mice lacking SynI and/or SynII display while CCK-GFP mice were kindly provided by Drs Steven A.
an epileptic phenotype and cognitive disabilities (Corradi et al. Siegelbaum and Peter Jonas. All experiments were performed
2008; Etholm et al. 2011, 2012, 2013; Michetti et al. 2017). in accordance with the guidelines established by the European
The deletion of SynI reduces the size of the reserve pool Communities Council (Directive 2010/63/EU of 22 September
(RP) of SVs alters short-term plasticity in excitatory synapses 2010) and approved by the Italian Ministry of Health.
(Rosahl et al. 1995; for review Cesca et al. 2010; Fassio et al.
2011b). In primary autaptic neurons, SynI deletion increases the
Preparation of Slices
amplitude of evoked excitatory postsynaptic currents (eEPSCs),
an effect that is attributable, at least in part, to a decreased Mice were anesthetized with an overdose of isoflurane, quickly
activation of presynaptic GABAB receptors (Chiappalone et decapitated and the brain was cut in horizontal slices of 400 μm
al. 2009; Valente et al. 2017). On the contrary, SynI knockout thickness using a Microm HM 650V microtome equipped with a
(KO) inhibitory synapses in primary neurons exhibit a decrease Microm CU 65 cooling unit (Thermo Fisher Scientific, Waltham,
in the amplitude of evoked inhibitory postsynaptic currents MA) at 2 ◦ C in a solution containing (in mM): 87 NaCl, 25 NaHCO3 ,
(eIPSCs) and a reduction of readily releasable pool [RRP]) (Baldelli 2.5 KCl, 0.5 CaCl2 , 7 MgCl2 , 25 glucose, 75 sucrose, and saturated
et al. 2007; Chiappalone et al. 2009). Interestingly, SynIIKO with 95% O2 and 5% CO2 . Slices were let to recover for 45 min at
inhibitory synapses of the dentate gyrus of the hippocampus 35 ◦ C and 30 min at room temperature (RT) in recording solution.
exhibit an increased eIPSC amplitude and RRP size associated
with the disappearance of asynchronous release, resulting
Electrophysiology
in a dramatic increase of the s/a-R and in a sharp decrease
of tonic inhibition (Medrihan et al. 2013). The latter result Whole-cell recordings were performed with a Multiclamp
was recently confirmed in inhibitory synapses of the CA1 700B/Digidata1440A system (Molecular Devices, Sunnyvale, CA)
region of the hippocampus, where SynII deletion increased the on visually identified mature dentate granule cells using a
s/a-R without affecting the eIPSC amplitude (Feliciano et al. BX51WI microscope (Olympus, Tokyo). All experiments were
2017). However, in this hippocampal region of SynIIKO mice, performed at RT, except for a subset of data, for which the
the increased s/a-R observed in regular spiking (presumably bath temperature was maintained at 35 ± 0.5 ◦ C thanks to
CCK/SOM) interneurons was associated with desynchronization an in-line solution heater in combination with a heated stage
of GABA release in fast-spiking (presumably PV) interneurons, chamber controlled by a feedback thermistor placed in the slice
suggesting distinct roles in various interneuron subpopulations chamber (Warner Instruments, Hamden, CT) The extracellular
(Feliciano et al. 2017). solution used for recordings contained (in mM): 125 NaCl,
Based on these findings, we aimed at clarifying whether 25 NaHCO3 , and 25 glucose, 2.5 KCl, 1.25 NaH2 PO4 , 2 CaCl2 ,
SynI has a specific role in controlling GABA release dynamics and 1 MgCl2 , pH 7.3 bubbled with 95% O2 –5% CO2 . A high-
in the various interneuron subpopulations. To this aim, we used chloride intracellular solution was used for all the experiments,
patch-clamp recordings of granule cells in the dentate gyrus of containing (in mM): 126 KCl, 4 NaCl, 1 MgSO4 , 0.02 CaCl2 , 0.1
wild-type (Wt) and SynIKO hippocampi. Opposite to what we BAPTA, 15 glucose, 5 HEPES, 3 ATP, and 0.1 GTP. pH 7.3 with
found in SynIIKO mice in the very same area (Medrihan et al. KOH, 290 mOsm/L. The patch pipette resistance was 3–5 MΩ
2013), we observed a strong reduction in synchronous release when filled with intracellular solution. All experiments were
and a concomitant increase in asynchronous release. We also performed in the presence of 50 μM D-APV, 10 μM CNQX, and
demonstrate that this effect was associated with a significantly 5 μM CGP55845 (all from Tocris) to block NMDA, AMPA, and
higher expression of SynI in PV interneurons than in CCK/SOM GABAB receptors, respectively. ω-Agatoxin IVA (AGA, 500 nM; The
interneurons. Optogenetic stimulation of specific interneuron Peptide Institute Inc.) was used to block P/Q-type Ca2+ channels.
populations revealed that SynI deletion impairs synchronous Tetrodotoxin (TTX, 0.3 μM; Tocris) was added to block AP induced
release in PV-interneurons and enhances asynchronous release by optogenetic activation of PV- and SOM-positive interneurons.
in SOM interneurons. The results demonstrate that SynI directly eIPSCs were recorded at a holding potential (Vh ) of −80 mV
contributes to the synchronization of GABA release and that the in response to extracellular stimulation of the molecular layer
expression of SynI/SynII can represent a push–pull mechanism with a monopolar glass electrode placed in the proximity of the
regulating the s/a-R of inhibitory transmission and thereby net- recorded granule cell. The minimal response was found at a
work excitability. stimulation intensity of 40 μA, lasting between 100 and 500 μs.
Synapsin I Synchronizes GABA Release Forte et al. 3

All data were acquired with Clampex and analyzed offline with all the interneurons contacting the patch-clamped granule cell.
Clampfit 10.2 (Molecular Devices, Sunnyvale CA, USA), Excel, Each stimulation step was followed by a high-frequency train
and GraphPad. (2 s @ 40 Hz) to evoke asynchronous release (Lu and Trussell
2000; Hagler and Goda 2001; Atluri 2006; Manseau et al. 2010).

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


The charge corresponding to the first 25 ms of the eIPSC was
Viral Injections and Optogenetic Stimulation
used as an estimation of synchronous release to normalize asyn-
Virus injection was performed on PV-Cre, PV-Cre/SynIKO, SOM- chronous release. The estimation of the synchronous charge
cre, and SOM-Cre/SynIKO mice. Mice from postnatal day 21–28 in this time frame (25 ms) avoids any possible contamination
were anesthetized using isoflurane and placed in a stereotaxic from the asynchronous release component since a correlation
frame. An injection needle (Hamilton 7000, 1 μL syringe) was exists between the extent of asynchronous release and the decay
lowered in the proximity of dentate gyrus and 1 μL of the Cre- phase of the eIPSC (Isaacson and Walmsley 1995; Medrihan et
dependent AAV1-DIO-ChETA-eYFP viral vector (1012 vp/mL; Penn al. 2013; Iremonger and Bains 2016). The asynchronous release
Vector Core) was infused at a rate of 0.1 μL/min (injection coor- was measured as the charge in the 10 s following the train
dinates: anterior–posterior −2 mm from Bregma, lateral ±1 mm, (2s@40Hz) for the extracellular stimulation experiments and was
dorsal–ventral −2.40 mm). Mice were returned to their home normalized, for each individual neuron, to the first 25 ms charge
cage and allowed to recover at least 3 weeks after surgery. Optical of the mean eIPSC average of 10 eIPSCs recorded at the same
stimuli on hippocampal slices were generated using a Cairn stimulation intensity before applying the train. The temporal
opto-LED system or a PRIZMATIX UHP-Mic-LED (wavelength: 473 profile of the delayed asynchronous release after the train (at
nm; light power at the source: 50 mW) by passing the light beam 160 μA stimulation intensity) was calculated by dividing the
through the microscope optical system and conveying it to the postsynaptic current recordings following the train in ten 1-
slice through an ×20 water-immersion objective. The light power s area bins. Asynchronous release in optogenetic experiments
measured with a power meter at the level of the slice surface was analyzed as the charge in the 2 s following the light stimula-
was approximately 4 mW/mm2 . Light stimuli (1-ms long) were tion and normalized with respect to the 2 s-area of spontaneous
TTL-triggered with Clampex digital output signals. activity before the train (Medrihan et al. 2013). For the calculation
of the GABA tonic current, we voltage clamped the granule
neurons at −80 mV and, after 30 s/1 min of baseline, we acutely
Data Analysis
applied bicuculline (100 μM) and calculated the difference in the
Ten to twenty consecutive eIPSCs at 0.1 Hz were averaged holding current before and after the treatment. The amplitude
to calculate peak amplitude and kinetics of the eIPSC. We and frequency of spontaneous IPSCs (sIPSCs), recorded at −80
considered the peak amplitude as the difference between the mV in the absence of TTX, were analyzed in a timeframe of 6 min
baseline and the peak of the evoked postsynaptic current and before the application of bicuculline and used for the analysis of
the charge as the area under the postsynaptic current. The the GABAergic tonic current.
decay time was calculated fitting the deactivation response
with a mono-exponential equation. The size of the RRP
In situ Hybridization
and the probability of synchronous release (RRPsyn and Pr ,
respectively) were calculated using the cumulative amplitude PV-Tomato Red, SOM-Tomato Red, and CCK-GFP male mice
analysis (Baldelli et al. 2007; Chiappalone et al. 2009; Medrihan were deeply anesthetized with 20% urethane (0.1 mL/10 g
et al. 2013). The estimation of the RRPsyn is based on the body weight) and perfused transcardially with 0.1 M phosphate
assumption that there is a rapid decrease in the number buffer containing 4% paraformaldehyde (pH 7.4). Brains were
of readily releasable quanta before reaching a steady state postfixed overnight, cryoprotected in 30% sucrose, and frozen
phase limited by the constant recycling of SVs and that in optimal cutting temperature. Horizontal sections (14 μm
equilibrium occurs between release and vesicle replenishment, thick) were cut using a CM 3050S Cryostat (Leica Microsystems,
with Pr during the train approaching 1 (Schneggenburger Wetzlar, Germany) and collected on Superfrost slides. Before
et al. 1999). RRPsyn was determined by summing up the peak starting hybridization, PV-tomato, SOM-tomato, and CCK-GFP
amplitudes of eIPSC responses during a 2-s stimulation train @ fluorescent signals in the hippocampal dentate gyrus area
40 Hz. Lower stimulation frequencies under our experimental were acquired using an Eclipse epifluorescence microscope
conditions did not fulfill the above-mentioned requirements (Nikon, Tokyo, Japan). The antisense sequence designed to
of the cumulative amplitude analysis. The number of linearly target SynIa/SynIb messenger RNAs (mRNAs) (atgaactacct gcg-
distributed data points starting from the last one (i.e., from gcgccgc ctgtcggaca gcaacttcat ggccaatctg ccgaatgggt acatgacaga
the 80th eEPSC), identified by adopting an objective criterion, cctgcagcgc ccgcaaccgc ccccgccgcc tccctcggcc gccagccctg gggc-
were fitted by linear regression and back extrapolated to time cacgcc cggctccgcg acagcctctg ccgagagggc ctccacagct gctccagtgg
0. The intercept with the Y-axis yielded the RRPsyn and the cttctccagc agcccctagt cctgggtcct cggggggcgg cggcttcttc tcgtcgctgt
ratio between the amplitude of the first eIPSC and the RRPsyn ctaacgcggt caagcaaacc acagcagccg cagccgccac cttcagcgag
yielded the Pr. To calculate paired-pulse depression (PPD), two caggtgggcg gtggctctgg gggcgcaggc cgcgggggcg ccgccgccag ggt-
brief stimulation pulses were applied at interstimulus intervals gctgctg gtcatcgacg aaccgcacac cgactgggca aaatacttca aagggaa-
ranging between 10 ms and 4 s (10 sweeps for each trial) and the gaa gatccatgga gaaattgaca ttaaagtaga gcaagctgaa ttctctgatc)
ratio between the second and the first response (paired-pulse was labeled with digoxigenin (DIG) using the 3 -DIG label-
ratio [PPR]) was calculated. ing kit (Roche, Milano, Italy). Sections were permeabilized
All the experiments for the evaluation of synchronous and with radioimmunoprecipitation assay buffer. After blocking
asynchronous release were performed using increasing stimula- of positive charges with acetic anhydride (0.25% v/v) in
tion intensities. We started from the minimal stimulation (40 μA) triethanolamine buffer (100 mM triethanolamine, 0.002% acetic
and increased the stimulation intensity of 3- (120 μA) and 4-fold acid) and washing with phosphate buffered saline (PBS), slices
(160 μA) to progressively recruit more inhibitory fibers from were prehybridized with hybridization solution (50% formamide,
4 Cerebral Cortex, 2019, Vol. 00, No. 00

5× SSC, 5× Denhardts, 50 μg/mL Harring Sperm DNA, 250 (Fig. 1d). In the same slices, we also ascertained whether these
μg/mL Yeast tRNA) at 70 ◦ C in hybridization solution. At day changes were associated with alterations in GABA tonic inhi-
2, sections were washed with posthybridization buffer (50% bition, as reported for SynII KO slices (Medrihan et al. 2015)
formamide, 2× SSC, 0.1% Tween-20) and B1 buffer (10× Maleic or spontaneous synaptic activity. Interestingly, in contrast with

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Acid, 0.1% Tween-20) at RT. Thereafter, slices were blocked with SynII KO slices, no significant changes were observed in both
B2 buffer (B1 buffer containing 10% normal goat serum) for 1 the tonic current (Fig. 1e) and the amplitude/frequency of sIPSCs
h and incubated with alkaline phosphatase-conjugated anti- (Fig. 1f ) in SynIKO slices.
DIG antibody (1:2000; Roche) overnight at 4 ◦ C. At day 3, slices SynI has been implicated in short-term plasticity changes
were washed with buffer B1 and then with buffer B3 (100 mM (increase in paired-pulse facilitation) at excitatory synapses
TrisCl pH 9.5, 50 mM MgCl2 , 100 mM NaCl, and 0.1%Tween-20) from hippocampal SynIKO or triple Syn KO slices (Rosahl et
once for 30 min and incubated overnight at 4 ◦ C with nitro blue al. 1993; Farisello et al. 2013), while PPD was not affected in
tetrazolium/5-bromo-4-chloro-3-indolyl phosphate (NBT/BCIP; inhibitory synapses of SynIKO primary hippocampal cultures
1 tablet in 10 mL ddH2 O; Roche). The reaction was terminated (Baldelli et al. 2007). Consistently, when inhibitory synapses
by several washes with 0.1% Tween-20 in PBS. After Hoechst of the dentate gyrus were excited by two consecutive stimuli
staining (1:400 in PBS) for 10 min, slices were air-dried and administered at interpulse intervals ranging between 10 ms and
mounted with ProLong Gold (Invitrogen). All solutions were 4 s, no differences in the paired-pulse ratios of monosynaptic
prepared using diethyl pyrocarbonate water and the cryostat eIPSCs were found between Wt and SynIKO mice (Fig. 1g,h).
cleaned with RNase ZAP (Sigma). The specificity of the SynI
probe was assessed using SynIKO hippocampal slices. The
mRNA signals were acquired using the Nikon Eclipse Microscope
A Decrease in RRPsyn in SynIKO Slices Accounts for the
and the signals of the interneuron markers and of SynI mRNA Reduced IPSC Amplitude
were merged using Adobe Photoshop and analyzed with ImageJ. We next investigated the cellular basis of the observed decrease
in the eIPSC amplitude using the cumulative amplitude pro-
file analysis during a high-frequency stimulation train (Fig. 2a;
Statistical Analysis
2 s @ 40 Hz (Schneggenburger et al. 1999). We confirmed the
Data are reported as box plots, except for the graphs, in which smaller amplitude of the first eIPSC in the train of SynIKO
means ± SEM are shown for clarity. The box plots elements neurons (Fig. 2b) and observed no significant differences in the
are: center line, median (Q2); square symbol, mean; box lim- eIPSC depression dynamics during the train (Fig. 2c). To per-
its, 25th (Q1)-75th (Q3) percentiles; whisker length is deter- form quantal analysis, the size of RRPsyn was calculated using
mined by the outermost data points. Normal distribution was a back extrapolation to time 0 of the linear portion of the
assessed using D’Agostino–Pearson’s normality test. In the case cumulative amplitude curve assuming that the slow linear rise
of two experimental groups, the statistical analysis was per- is attributable to the equilibrium between the release-induced
formed using the two-tailed unpaired Student’s t-test for nor- depletion and the constant replenishment of the RRPsyn with
mally distributed data or the Mann–Whitney’s U-test for non- Pr approaching 1 (Fig. 2d). As previously shown in primary neu-
normally distributed data. One-way ANOVA/Holm–Sidak’s tests rons (Baldelli et al. 2007; Chiappalone et al. 2009), the analy-
or Kruskal–Wallis/Dunn’s tests were used to analyze more than sis clearly indicated that the decreased eIPSC amplitude was
two experimental groups with normally or non-normally dis- entirely attributable to a decline in RRPsyn , while Pr was not
tributed data, respectively. significantly affected (Fig. 2f ).

Results The Balance between Synchronous and Asynchronous


Deletion of SynI Reduces the Amplitude of eIPSCs Release of GABA Is Altered in SynIKO Mice
without Affecting Short-Term Plasticity
We previously showed that SynII deletion is associated with a
We previously reported that SynII deletion promotes upregu- total loss of asynchronous GABA release at inhibitory synapses
lation of synchronous GABA release and a concomitant loss of the hippocampal dentate gyrus (Medrihan et al. 2013). Thus,
of delayed asynchronous release at GABAergic synapses of the we investigated whether also SynI deletion has an impact on the
hippocampal dentate gyrus (Medrihan et al. 2013, 2015). Since s/a-R of GABA at the same synapses. We stimulated presynaptic
we previously showed that SynI deletion impairs GABAergic fibers by a single pulse of increasing intensity at RT. A weak
transmission in primary hippocampal cultures (Baldelli et al. stimulus (40 μA) was set to a value capable to activate the
2007; Chiappalone et al. 2009), we investigated whether it has smallest postsynaptic current, putatively due to the activation
similar or distinct functions on synchronous and asynchronous of a single presynaptic inhibitory fiber. Subsequently, we pro-
release with respect to SynII. To this aim, we performed whole- gressively increased the stimulation intensity (up to 160 μA)
cell voltage-clamp recordings of dentate gyrus granule neurons to recruit a larger number of inhibitory fibers and yield an
in 3–5 weeks old SynIKO and Wt mice, a period in which SynIKO average estimation of the total amount of synchronous and
mice are still nonepileptic (Rosahl et al. 1995; Corradi et al. 2008). asynchronous release impinging onto the patch-clamped gran-
We placed an extracellular electrode in the molecular layer to ule cell. For each stimulation intensity, single pulses (each last-
evoke the smallest eIPSC, putatively due to the activation of ing 100–500 μs) were applied at 0.1 Hz over 1 min. To quantify
a single presynaptic fiber onto granule cells (Medrihan et al. the amount of asynchronous release we used high-frequency
2013). A significant reduction of eIPSC amplitude was observed stimulation trains (2 s train @ 40 Hz) that evoke a delayed release
in SynIKO slices with respect to Wt slices (Fig. 1a,b). This effect that persists for hundreds of milliseconds to seconds after the
was paralleled by a reduction in the total eIPSC area (Fig. 1c), end of the stimulation (Lu and Trussell 2000; Hagler and Goda
in the absence of significant changes in the current kinetics 2001; Farisello et al. 2013).
Synapsin I Synchronizes GABA Release Forte et al. 5

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019

Figure 1. SynI deletion reduces the amplitude of the eIPSC without affecting PPD. (a) Representative traces of eIPSCs from Wt and SynIKO dentate gyrus granule
neurons. (b–d) Box plots of eIPSC amplitude (b; 49.88 ± 8.25 pA; n = 28 neurons from ten mice for Wt neurons; 34.93 ± 4.07 pA; n = 28 neurons from six mice for
SynIKO neurons; ∗ P < 0.05, unpaired Student’s t-test with Welch’s correction), charge (c; median 1142 and 643.7 (pA × ms) × 103 for Wt and SynIKO respectively;
∗∗ P < 0.01, Mann–Whitney U-test) and decay time (d; 27.97 ± 1.94 ms and 22.53 ± 2.02 ms for Wt and SynIKO respectively; P = 0.57, unpaired Student’s t-test with
Welch’s correction). (e) Representative traces of tonic inhibitory currents (left) and box plot of GABA tonic inhibition (right) in Wt and SynIKO granule neurons (P = 0.15,
unpaired Student’s t-test with Welch’s correction; n=11 neurons from four Wt and n = 10 from three SynIKO mice). (f ) Representative traces of spontaneous inhibitory
currents (sIPSC; left) recorded in the absence of TTX and box plots of sIPSC amplitude (middle) and frequency (right) in Wt and SynIKO neurons (sIPSC amplitude: P =
0.98, unpaired Student’s t-test with Welch’s correction; sIPSC frequency: P = 0.14, Mann–Whitney U-test; n = 11 neurons from four Wt and n = 10 from three SynIKO
mice). (g) Representative dual eIPSCs traces recorded from dentate gyrus Wt and KO neurons in response to paired stimulation of the inhibitory fibers at the indicated
interstimulus intervals. (h) Mean (± SEM) PPR observed in Wt and KO neurons plotted as a function of the interevent interval (n = 18 neurons from six Wt mice and n
= 19 neurons from seven SynIKO mice; P > 0.1 at all intervals, Student’s t-test with Welch’s correction).
6 Cerebral Cortex, 2019, Vol. 00, No. 00

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Figure 2. The effect of SynI deletion on eEPSC amplitude is due to a decrease of the RRPsyn . (a) Representative responses from Wt and SynIKO neurons induced by
stimulation of the inhibitory fibers in the molecular layer of dentate gyrus for 2 s @ 40 Hz. Stimulation artifacts were removed for clarity. (b) Box plot of the amplitude
of the first eIPSC in the train (median 114.90 pA; n = 13 neurons from seven mice for Wt; median 60.71 n = 15 neurons from five mice for SynIKO; ∗∗ P < 0.01, Mann–
Whitney U-test). (c) The plot of the normalized amplitude (means ± SEM) versus time showing the multiple-pulse depression during the 2 s @ 40 Hz train in Wt and KO
neurons. (d) Cumulative amplitude profile of a 2 s train @ 40 Hz. Data points of the last 30 responses were fitted by linear regression and the line was extrapolated to
time 0 to estimate the RRPsyn . (e) Box plot of RRPsyn in Wt and SynIKO mice (1017 ± 152 pA and 588.80 pA ± 77.58 pA for Wt and SynIKO mice respectively; ∗ P < 0.05,
unpaired Student’s t-test with Welch’s correction). (f ) Box plot of the Pr in Wt and SynIKO mice (0.162 ± 0.02 and 0.144 ± 0.020 for Wt and SynIKO mice respectively; P
= 0.52, unpaired Student’s t-test with Welch’s correction).

To yield an estimation of the synchronous charge not con- GABA release (Fig. 4b–d), were preserved in SynIKO slices at 35 ◦ C.
taminated by any asynchronous component that may affect Moreover, under these conditions, the increase in the asyn-
the decay phase of the eIPSC, we referred to the charge cor- chronous release was significant even immediately after the end
responding to the first 25 ms of the eIPSC (Fig. 3a, left panel). of the train ∼5 s after the high-frequency stimulation (Fig. 4e).
The reliability of this estimation was shown by the linear cor- These data indicate the existence of nonredundant functions of
relation found between the RRPsyn and the respective 25 ms Syns in the regulation of GABA release dynamics.
area of the first eIPSC in the train (Fig. 3a, middle panel). The
synchronous charge transferred by GABA release was markedly
lower in SynIKO than in Wt slices at all stimulation intensi-
Blockade of P/Q Ca2+ Channels Selectively Occludes
ties (Fig. 3a, right panel). On the contrary, the absolute charge the Impairment in Synchronous GABA Release Due
transferred by asynchronous release was 2-fold larger in SynIKO to SynI Deletion
than in Wt slices at 160 μA stimulation intensity (Fig. 3b,c). When It is widely accepted that synchronous synaptic transmission
the asynchronous release was normalized to the 25-ms syn- mainly relies on the AP-induced activation of P/Q-type Ca2+
chronous charge, that is, to a parameter reflecting the num- channels and, consistently, synchronous GABA release is largely
ber of activated presynaptic fibers and the quantal parameters inhibited by 500 nM AGA, a selective P/Q-type channels antago-
of release, its normalized values were strongly increased in nist (Hefft and Jonas 2005).
SynIKO slices at all stimulation intensities (Fig. 3d). The time- To evaluate the contribution of P/Q-type Ca2+ channels
course of asynchronous release in SynIKO synapses confirmed to the imbalance of synchronous/asynchronous inhibitory
the enhanced release from 2 to 5 s after the high-frequency transmission of SynIKO mice, we repeated the experiments in
train and was characterized by a much slower return to baseline the presence of AGA (500 nM). In Wt slices, the synchronous
than in Wt slices (Fig. 3e). The effect of SynI deletion was even release elicited by 160 μA extracellular stimulation was
enhanced when the asynchronous release was normalized to markedly reduced in the presence of AGA, while the AGA effect
the area of sIPSCs recorded before the train (not shown). was strongly attenuated in SynIKO slices (Fig. 5a,b). On the
Since it was shown that, in excitatory synapses, increasing contrary, AGA did not affect the extent of asynchronous release
the temperature to physiological levels decreases asynchronous in both genotypes (Fig. 5c). From the above data, we calculated
release (Pyott and Rosenmund 2002), we repeated the above- the contribution of P/Q-type Ca2+ channels to synchronous and
described experiments at 35 ◦ C (Fig. 4). Both, the decrease of asynchronous release in the two genotypes by subtracting the
synchronous release (Fig. 4a) and the increase in asynchronous median current recorded in presence of AGA from that recorded
Synapsin I Synchronizes GABA Release Forte et al. 7

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Figure 3. The balance between synchronous and asynchronous GABA release is altered in SynIKO mice. (a) Left: Representative traces of eIPSCs at 160 μA of stimulation;
the gray 25-ms area of the eIPSC was taken as a purely synchronous fraction of the eIPSC charge used to normalize the asynchronous charge. Center: correlation between
the charge of the first 25 ms of eIPSCs in the train and the respective RRPsyn calculated using cumulative amplitude analysis. Pearson’s correlation coefficient, ∗∗ P =
0.04. Right: Box plots of the first 25-ms fraction of synchronous charge for eIPSCs evoked at increasing stimulation amplitudes in Wt (black symbols) and SynIKO (gray
symbols) mice. (b) Representative traces of asynchronous GABA release (gray areas) after a 2-s train stimulation (160 μA; 40 Hz) in Wt and SynIKO mice; the gray vertical
lines represent the ten 1-s bins used to evaluate the time course of the asynchronous charge in the two genotypes. (c) Box plots of asynchronous charge calculated in
the first 10 s after the train evoked at increasing stimulation amplitudes in Wt and SynIKO mice. (d) Box plots of asynchronous release as a function of the increasing
amplitude of stimulation and normalized to the 25-ms fraction of synchronous release. (e) Time course of asynchronous charge (mean absolute values ± SEM) after
the end of the stimulation train. For each experiment, at least 10 neurons from three mice per genotype were recorded. ∗ P < 0.05, ∗∗ P < 0.01, unpaired Student’s t-test
with Welch’s correction.

under control conditions (Fig. 5d,e). This analysis demonstrates Synchronous GABA Release in PV Interneurons Is
that the effect of AGA on synchronous release is occluded in Strictly Dependent on Syn I Expression
SynIKO slices and that the synchronous release impairment due
to SynI deletion largely corresponds to the AGA-sensitive, P/Q It was previously shown that the dynamics of synchronous
Ca2+ channel-mediated fraction of GABA release (Fig. 5d). On GABA release from PV interneurons relies on the recruitment of
the contrary, AGA-sensitive asynchronous release is virtually P/Q-type Ca2+ channels (Hefft and Jonas 2005). The observation
lacking in both genotypes, demonstrating its full independence that P/Q Ca2+ channel blockade occludes the SynIKO-induced
from P/Q-type Ca2+ channels (Fig. 5e). reduction in synchronous GABA release led us to hypothesize
8 Cerebral Cortex, 2019, Vol. 00, No. 00

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Figure 4. The alteration of synchronous and asynchronous GABA release in SynIKO mice is preserved at physiological temperature. (a) Left: Representative traces of
eIPSC at minimal stimulation at 35 ◦ C; the gray 25-ms area of the eIPSC was taken as a purely synchronous fraction of the eIPSC charge used to normalize asynchronous
charge. Right: Box plots of the first 25-ms fraction of synchronous charge for eIPSCs evoked in Wt (black symbols) and SynIKO (gray symbols) mice. (b) Representative
traces of asynchronous GABA release (clear gray areas) after a 2-s train stimulation @ 40 Hz in Wt and SynIKO mice; the gray vertical lines define the 1-s bins used to
evaluate the time course of the asynchronous charge in the two genotypes. (c) Box plots of asynchronous charge calculated in the first 10 s after a 2-s train @ 40 Hz.
(d) Box plots of asynchronous release normalized to the 25-ms fraction of synchronous release. (e) Time course of asynchronous charge (mean absolute values ± SEM)
plotted as a function of time after the end of the stimulation train. For each experiment n = 13 neurons from three mice per genotype were recorded. ∗ P < 0.05, ∗∗ P <
0.01, unpaired Student’s t-test with Welch’s correction.

that it may specifically involve PV interneurons. Thus, we light-dependent activation of PV/SynIKO synapses onto granule
crossed SynIKO mice with PV-cre mice and successively injected cells phenocopied the reduction in amplitude of synchronous
a Cre-dependent AAV viral vector in the dentate gyrus to GABA release evoked by electrical stimulation (Fig. 6c,e), while no
selectively express the fast excitatory opsin ChETA in PV changes were observed in the amount of asynchronous release
neurons (Fig. 6a). We then recorded IPSCs induced by single after high-frequency trains of light stimuli (Fig. 6f ).
light pulses (at 470 nm) or trains of light stimuli (2 s @ 40 Hz)
in patched granule cells of the dentate gyrus (Fig. 6b,d). The
application of TTX inhibited the release of GABA from PV
SynI Is Predominantly Expressed in PV Interneurons of
interneurons, demonstrating that light-evoked GABA release the Dentate Gyrus
was strictly dependent on the AP generation, and not on local Distinct interneuron subpopulations display the propensity for
depolarization of the presynaptic site (Fig. 6b). The selective either synchronous or asynchronous release and are therefore
Synapsin I Synchronizes GABA Release Forte et al. 9

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Figure 5. ω-Agatoxin IVA occludes the effect of SynI deletion on synchronous GABA release. (a) Representative traces of eIPSCs from Wt (left) and SynIKO (right) slices
incubated in the absence or presence of the P/Q type Ca2+ channel inhibitor ω-agatoxin IVA (AGA; 500 nM). (b) Box plots of the fraction of synchronous release in the
four experimental groups, calculated in the first 25 ms of the synaptic response at 160 μA stimulation intensity (42.77 ± 4.89 pA × ms; n = 9 neurons from five mice
and 9.63 ± 1.63 pA × ms; n = 8 neurons from three mice for Wt and Wt/AGA groups, respectively; 22.86 ± 2.43 pA × ms; n = 10 neurons from three mice and 12.92 ±
2 pA × ms n = 9 neurons from three mice for SynIKO and SynIKO/AGA groups, respectively). ∗ P < 0.05, ∗∗∗ P < 0.001, one-way ANOVA/Holm–Sidak’s tests. (c) Box plots
of the asynchronous release calculated in the first 10 s after a 2s-train @ 40 Hz evoked at 160 μA stimulation intensity in the four experimental groups (6.90 ± 1.33
and 6.72 ± 1.34 pA × ms for Wt and Wt/AGA groups, respectively; 13.49 ± 3.09 and 15.42 ± 2.07 pA × ms; for SynIKO and SynIKO/AGA groups, respectively). ∗ P < 0.05,
one-way ANOVA/Holm–Sidak’s tests. (d,e) AGA-sensitive (black) and AGA-insensitive (gray) fractions of synchronous (d) and asynchronous (e) release. The percentages
of synchronous and asynchronous charge are shown.

characterized by different s/a-Rs. In PV interneurons, GABA terminals, we measured the codistribution of SynI mRNAs with
release is more temporally precise than in CCK/SOM interneu- the specific markers (PV, CCK, and SOM) of the three main
rons. Indeed, eIPSCs generated by CCK/SOM interneurons are GABAergic neuronal subpopulations in the dentate gyrus by
characterized by long delays, strong fluctuations in latency in situ hybridization (Fig. 7a). Interestingly, SynI mRNA was
and amplitude as well as high failure rates (∼16%), thereby expressed at significantly higher levels in "synchronous" PV
generating an unstable inhibitory output signal (Hefft and Jonas interneurons than in “asynchronous” CCK/SOM interneurons
2005; Daw et al. 2009; Ali and Todorova 2010; Savanthrapadian (Fig. 7b).
et al. 2014; Yavorska and Wehr 2016; Matos et al. 2018).
Since it was demonstrated in cell cultures that hippocampal
Deletion of SynI Increases Asynchronous Release
interneurons collectively express all Syn isoforms (Song and
in SOM Interneurons
Augustine 2016), we evaluated the possibility that SynI and
SynII are differentially expressed in specific interneuron The selective decrease of synchronous release in PV interneu-
subpopulations. Being Syns strictly localized to synaptic rons, together with the differential expression of SynI,
10 Cerebral Cortex, 2019, Vol. 00, No. 00

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Figure 6. Synapsin I KO selectively reduces the synchronous release in PV interneurons. (a) Example of a hippocampal slice from PV-cre mice infected with AAV1-DIO-
ChETA-eYFP Cre-dependent viral vector; GC (Granule Cell Layer). (b) Representative traces of a single eIPSC (left) evoked by 470 nM blue light pulses in PV/Wt (black
trace) and PV/SynIKO slices (red trace). The application of TTX fully inhibited light-evoked GABA release (lower trace). (c) Box plot of eIPSCs amplitude (162.7 ± 21 pA, n
= 15 from four PV/Wt mice; 90.36 ± 14.16 pA, n = 16 from six PV/SynIKO mice). ∗∗ P < 0.01, unpaired Student’s t-test with Welch’s correction). (d) Representative traces
of inhibitory currents evoked by a 2 s-train of light pulses @ 40 Hz evoked in PV/Wt (black trace) and PV/SynIKO (red trace) slices. (e) Box plot of synchronous release
(18.14 ± 1.96 pA × ms × 102 , n = 15 from four PV/Wt mice; 8.6 ± 0.93 pA × ms × 102 , n = 16 from six PV/SynIKO mice; ∗∗∗ P < 0.001; Mann–Whitney U-test) from PV
interneuron of Wt and SynIKO mice calculated in the first 25 ms of the first eIPSC of the 2 s-train @ 40 Hz. (f ) Mean (±SEM) asynchronous release normalized to the 2
s-pretrain spontaneous release charge and plotted against time after the end of the stimulation train (P ≈ 0.5 for each point, Mann–Whitney U-test, n = 10 from five
PV/Wt mice; n = 6 from five PV/SynIKO).

suggests that SynI plays a fundamental role in the expres- neurotransmitter release is an essential factor in converting the
sion of synchronous GABA release, constraining the natural digital signaling of the AP into the analog process of synaptic
tendency of neurons to release GABA asynchronously. If this transmission. The synchronous release is at the basis of the
hypothesis holds true, the deletion of SynI from asynchronous cellular communication in the brain; it is fundamental for
interneurons should increase the latter release modality. all cerebral functions as demonstrated by the early death of
Thus, we investigated the specific effects of SynI deletion in synaptotagmin I KO mice in which the synchronous release is
these interneurons by generating SOM-cre/SynIKO mice and completely disrupted (Geppert et al. 1994).
expressing the fast channelrhodopsin ChETA specifically in The synchronous release is a hallmark of PV-interneurons
SOM interneurons by in vivo transduction of dentate gyrus expressing P/Q Ca2+ channels, while asynchrony of release char-
with the AAV1-DIO-ChETA-eYFP viral vector. Deletion of SynI acterizes the other interneuronal populations including CCK
had no effects on the very low levels of synchronous release interneurons predominantly expressing N-type Ca2+ channels
evoked in these neurons by single light stimuli, when either and SOM interneurons (Hefft and Jonas 2005; Savanthrapadian
amplitude or transferred charge was considered (Fig. 7c,d). et al. 2014). The time-locking of GABA release with the AP
However, when asynchronous GABA release was elicited by a 2 (synchronous release) achieves a point-to-point space locking
s-train of light stimuli @ 40 Hz, a significant increase in the early of phasic inhibition signal, while asynchronous GABA release
poststimulation phase was observed at SOM/SynIKO synapses is integrated in time and represents a more tonic signal that,
(Fig. 7e,f ). followed by spillover, generates a tonic inhibition that spreads
in the extracellular volume (volume control of excitability). A large
body of experimental evidence supports the view that syn-
Discussion chronous GABA release by PV interneurons in the hippocampus
Synchronous and asynchronous modalities of GABA release plays important roles in network physiology such as feedback
play an important role in the regulation of network activity and feedforward inhibition, oscillations, pattern identification,
and excitability. Modifications in the Ca2+ wave dynamics, Ca2+ temporal coding and information processing, network plasticity
sensitivity or spectrum of molecular actors participating in and learning (Bartos et al. 2002; Hu et al. 2014; Gan et al. 2017;
the neuronal communication lead to substantial changes in Espinoza et al. 2018).
synaptic transmission that can result in pathological conditions, Although the Ca2+ transient following a single AP disappears
such as epilepsy (Kaeser and Regehr 2014). The timing of after few milliseconds, sustained firing activity gives rise to
Synapsin I Synchronizes GABA Release Forte et al. 11

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Figure 7. SynI is predominantly expressed in PV interneurons and increases the asynchronous release in SOM interneurons. (a) Representative images of the DG region
in brain slices from PV-tomato, CCK-GFP and SOM-tomato mice, respectively (red). SynI mRNA, analyzed by in situ hybridization, is visualized in green. Scale bar,
100 μm. (b) Box plots of the SynI mRNA expression in PV, CCK and SOM interneurons (61, 41 and 100 neurons from three PV-tomato, CCK-GFP and SOM-tomato mice
respectively). ∗∗ P < 0.01, ∗∗∗ P < 0.001, Kruskal–Wallis/Dunn’s tests. (c) Representative traces of a single eIPSCs evoked by 470 nM blue light pulses in SOM (black trace)
and SOM/SynIKO (red trace) slices. (d) Peak amplitude (SOM/Wt: 13.16 ± 3.78 pA, mean ± SEM, n = 7; SOM/SynIKO: 20.62 ± 5.85 pA, n = 8, from three animals per group;
P = 0.32) and area (SOM/Wt: 2.175 ± 0.6827 pA × ms × 102 , mean ± SEM, n = 7; SOM-SynIKO: 3.61 ± 1.191 pA × ms × 102 , mean ± SEM, n = 8; from three animals
per group; P = 0.33, Mann–Whitney U-test) of synchronous GABA release in SOM interneurons were unaffected by the deletion of SynI. (e) Representative traces of
inhibitory currents evoked by a 2 s-train of light pulses @ 40 Hz evoked in SOM/Wt (black trace) and SOM/SynIKO (red trace) slices. (f ) Mean (±SEM) asynchronous
release normalized to 2 s-pretrain spontaneous release charge and plotted against time after the end of the stimulation train. (SOM/Wt n = 6 neurons and SOM/SynIKO
n = 5 neurons from three mice per experimental group; ∗ P < 0.05, Mann–Whitney U-test).

an increase in the amount of the residual nonbuffered Ca2+ , integrated over time, it is also causally related to the genera-
stimulating release in a desynchronized fashion with respect tion and extent of tonic shunting inhibition (Medrihan et al.
to the AP (Hagler and Goda 2001; Otsu 2004). Asynchronous 2013). Because of this, the asynchronous release was found to
release is a relevant component of neuronal communication in be involved in neurological disorders, such as epilepsy, spinal
specific areas of the central nervous system, such as the spinal muscular atrophy, or Alzheimer’s disease (Yang et al. 2007; Ruiz
cord dorsal horn, deep cerebellar nuclei, the inferior olive or et al. 2010; Jiang et al. 2012)
specific interneuron subtypes present in all brain areas (Hefft Syns are important regulators of SV trafficking and thereby
and Jonas 2005; Best and Regehr 2009; Labrakakis et al. 2009; Ali participate in regulating the equilibrium between RRP and RP in
and Todorova 2010; Savanthrapadian et al. 2014). Asynchronous nerve terminals (Cesca et al. 2010). Deletion of either Syn gene
GABA release could generate a fluctuating and long-lasting causes an epileptic propensity that phenotypically emerges at
inhibitory signal ideally suited for the purpose of gain control 2–3 months of age in KO mice, in parallel with the progressive
in postsynaptic target cells (Mitchell and Silver 2003) and being postnatal increase in the expression of Syns I/II at synapses
12 Cerebral Cortex, 2019, Vol. 00, No. 00

(Bogen et al. 2009). This has led to the idea that Syn I and II may Our results indicate that SynI can be added to the task
have redundant functions, given also the sequence identity in force of synaptic proteins favoring GABA synchronous release;
their large N-terminal region. We previously described that SynII its expression, particularly prominent in PV-interneurons, is
is essential for asynchronous GABA release to occur, and the essential to assure an efficient phasic release of GABA in these

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


specific deficit of asynchronous GABA release in SynIIKO mice neurons. In addition to boosting synchronous release, SynI could
promotes a marked decrease in tonic inhibition, resulting in also act as an inhibitory constraint for asynchronous release.
hyperexcitability and epileptogenesis. A recent paper confirmed Indeed, SOM/CCK neurons constitutively expressing low SynI
this result and showed, using specific Ca2+ channel blockers, levels display a prevalent asynchronous release that is further
that SynII deletion suppressed the asynchronous release com- increased by the genetic deletion of SynI. It is also possible that
ponent at synapses dependent on N-type, but not P/Q-type, Ca2+ low levels or absence of SynI, by simply removing a positive
channels (Feliciano et al. 2017). effect on synchronous release, allow the asynchronous machin-
SynI, in addition to its major role in regulating the entry/exit ery to take over.
of SVs into/from the RP, was shown to play a postdocking role in The precise molecular mechanisms, by which SynI boosts
facilitating the last obligatory steps of exocytosis in primary hip- synchronous GABA release of PV interneurons and asyn-
pocampal neurons, an effect that was specific for inhibitory neu- chronous GABA release of SOM interneurons, remains to be
rons (Baldelli et al. 2007; Chiappalone et al. 2009; Lignani et al. identified. We have demonstrated that SynI is critical for
2013). Based on these results, we wanted to clarify whether SynI maintaining the correct size of RRPsyn in inhibitory synapses
has similar or distinct roles in the dynamics of GABA release (Baldelli et al. 2007; Chiappalone et al. 2009). In PV-positive
with respect to SynII in the intact circuits of the dentate gyrus terminals, RRPsyn is made of SVs docked and primed within
of the hippocampus. We have found that, in SynIKO mice, syn- the “nanodomains” of P/Q-type Ca2+ channels (Hefft and Jonas
chronous GABA release is impaired due to a reduction of RRPsyn 2005). The lack of SynI is likely to increase the mobility of the
size, while the asynchronous release is significantly increased. docked/primed SVs that can escape the very small “nanodomain”
Notably, these changes in release dynamics are opposite to around the P/Q-type Ca2+ channels, thus decreasing RRPsyn .
what found in the very same synapses lacking SynII, in which This hypothesis suggests that the smaller RRPtot measured by
synchronous release is increased and asynchronous release is hypertonic stimulation in SynIKO autaptic inhibitory neurons
virtually abolished (Medrihan et al. 2013). The potentiation of (Baldelli et al., 2007) could be due to the high frequency of PV
synchronous release by SynI is occluded by blockade of P/Q- interneurons in cultured autaptic neurons from the hippocam-
type Ca2+ channels, confirming that it acts at SVs located in the pus. Indeed, in PV-positive terminals, the SVs leaving the RRPsyn
nanodomain associated with these active zone-targeted chan- cannot populate the asynchronous release pool because of the
nels. Interestingly, SynI was highly expressed in PV interneu- short lasting and spatially confined Ca2+ -transients resulting
rons characterized by synchronous release and P/Q Ca2+ chan- from the strong Ca2+ buffering action of PV. The enhanced SV
nels, while lower SynI levels characterize CCK/SOM interneu- mobility in the absence of SynI could also explain the increased
rons expressing asynchronous activity and non-P/Q Ca2+ chan- asynchronous release observed in SOM/CCK interneurons. In
nels. Selective optogenetic stimulation of SynI depleted PV- these neurons, the higher SV mobility increases the probability
interneurons was associated with a marked impairment of syn- that SVs enter the large “microdomains” surrounding N-type
chronous GABA release that was paralleled by an increased channels that characterize the asynchronous release dynamics
asynchronous GABA release from SynI depleted SOM interneu- in CCK/SOM terminals devoid of presynaptic Ca2+ buffers.
rons that phenocopied the dual effect observed with electrical The deletion of SynI did not apparently affect the small
stimulation. fraction of synchronous release activity displayed by CCK/SOM
Several studies were performed to uncover the molecular interneurons. If SynI is required by P/Q-type Ca2+ channels
actors orchestrating the balance between synchronous and to operate synchronous release in PV-positive terminals, the
asynchronous release in neurons. A task force of presynaptic lack of these channels, correlating with a release that occurs
proteins favoring synchronous release exists, comprising P/Q- within “microdomains,” in CCK/SOM terminals, together with
type Ca2+ channels, synaptotagmins I/II (Nishiki and Augustine the lower SynI expression in these neurons, may explain the
2004; Bornschein and Schmidt 2018; Chang et al. 2018) VAMP2 lack of genotype effect on synchronous release in CCK/SOM
(Maximov et al. 2009), Ca2+ buffering proteins (e.g., PV) (Hefft and interneurons.
Jonas 2005) counterbalancing the pool of presynaptic proteins In conclusion, our data indicate that SynI and SynII represent
favoring asynchronous release, whose members include N- a system for fueling both phasic and tonic GABA release
type Ca2+ channels (Hefft and Jonas, 2005), synaptotagmin and, although they play a distinct role for the two release
VII (Chen and Jonas 2017; Chen et al. 2017), Doc2 (Yao et al. modalities, they both contribute to the inhibitory control of
2011), VAMP4 (Raingo et al. 2012), and SynII (Medrihan et al. network excitability and are causative genes for epileptic
2013, 2015; Bornschein and Schmidt 2018). These two pools synaptopathies.
of presynaptic proteins likely act as a push–pull mechanism
controlling the timeliness of neurotransmitter release with
respect to AP pacing. This dynamic control is strongly affected
Funding
by the differential expression of these proteins in the particular
neuronal subpopulation, thereby generating groups of neurons, Compagnia di San Paolo Torino (grant IDs ROL 20612 and 9344);
in which the synchronous or asynchronous release modality Ministero della Salute Ricerca Finalizzata (GR-2016-02363972);
predominates. This is the case of the main hippocampal PV- EU Era-Net Neuron 2017 “Snareopathies” and ITN ECMED
positive interneurons that are mostly synchronous and of (Grant agreement no. 642881); Italian Ministry of University
SOM/CCK interneurons that release GABA in a preferentially and Research (PRIN2017 Immune-synaptopathies: dissecting
asynchronous fashion. the contribution of inflammation to synaptic dysfunctions).
Synapsin I Synchronizes GABA Release Forte et al. 13

Notes Corradi A, Zanardi A, Giacomini C, Onofri F, Valtorta F, Zoli


M, Benfenati F. 2008. Synapsin-I- and synapsin-II-null mice
We thank Drs Steven A. Siegelbaum (Howard Hughes Medical
display an increased age-dependent cognitive impairment. J
Institute, Columbia University Medical Center, New York State
Cell Sci. 121:3042–3051.
Psychiatric Institute, New York, NY) and Peter Jonas (Institute

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


Daw MI, Tricoire L, Erdelyi F, Szabo G, McBain CJ. 2009.
of Science and Technology, Vienna, Austria) for kindly providing
Asynchronous transmitter release from cholecystokinin-
the CCK-GFP mice. We also thank Drs Caterina Michetti, Mon-
containing inhibitory interneurons is widespread and target-
ica Morini (Italian Institute of Technology, Genova, Italy) and
cell independent. J Neurosci Off J Soc Neurosci. 29:11112–11122.
Michele Cilli (IRCCS San Martino, Genova, Italy) for help in breed-
Espinoza C, Guzman SJ, Zhang X, Jonas P. 2018. Parvalbumin+
ing the mice, Silvia Casagrande (Department of Experimental
interneurons obey unique connectivity rules and establish a
Medicine, University of Genova, Italy), Arta Mehilli and Diego
powerful lateral-inhibition microcircuit in dentate gyrus. Nat
Moruzzo (Center for Synaptic Neuroscience, Istituto Italiano di
Commun. 9:4605.
Tecnologia, Genova, Italy) for assistance in genotyping assays
Etholm L, Bahonjic E, Heggelund P. 2013. Sensitive and critical
and in the preparation of primary cultures.
periods in the development of handling induced seizures in
Conflict of Interest: None declared.
mice lacking synapsins: differences between synapsin I and
synapsin II knockouts. Exp Neurol. 247:59–65.
References Etholm L, Bahonjic E, Walaas SI, Kao H-T, Heggelund P. 2012. Neu-
Ali AB, Todorova M. 2010. Asynchronous release of GABA via roethologically delineated differences in the seizure behavior
tonic cannabinoid receptor activation at identified interneu- of synapsin 1 and synapsin 2 knock-out mice. Epilepsy Res.
ron synapses in rat CA1. Eur J Neurosci. 31:1196–1207. 99:252–259.
Atluri PP. 2006. The kinetics of synaptic vesicle Reacidification Etholm L, Lindén H, Eken T, Heggelund P. 2011. Electroencephalo-
at hippocampal nerve terminals. J Neurosci. 26:2313–2320. graphic characterization of seizure activity in the synapsin
Baldelli P, Fassio A, Valtorta F, Benfenati F. 2007. Lack of synapsin I/II double knockout mouse. Brain Res. 1383:270–288.
I reduces the readily releasable pool of synaptic vesicles Farisello P, Boido D, Nieus T, Medrihan L, Cesca F, Valtorta F,
at central inhibitory synapses. J Neurosci Off J Soc Neurosci. Baldelli P, Benfenati F. 2013. Synaptic and extrasynaptic ori-
27:13520–13531. gin of the excitation/inhibition imbalance in the hippocam-
Bartos M, Vida I, Frotscher M, Meyer A, Monyer H, Geiger JRP, pus of synapsin I/II/III knockout mice. Cereb Cortex N Y N. 1991,
Jonas P. 2002. Fast synaptic inhibition promotes synchronized 23:581–593.
gamma oscillations in hippocampal interneuron networks. Fassio A, Patry L, Congia S, Onofri F, Piton A, Gauthier J, Pozzi
Proc Natl Acad Sci U S A. 99:13222–13227. D, Messa M, Defranchi E, Fadda M et al. 2011a. SYN1 loss-
Best AR, Regehr WG. 2009. Inhibitory regulation of electrically of-function mutations in autism and partial epilepsy
coupled neurons in the inferior olive is mediated by asyn- cause impaired synaptic function. Hum Mol Genet.
chronous release of GABA. Neuron. 62:555–565. 20:2297–2307.
Bogen IL, Jensen V, Hvalby O, Walaas SI. 2009. Synapsin- Fassio A, Raimondi A, Lignani G, Benfenati F, Baldelli P. 2011b.
dependent development of glutamatergic synaptic vesicles Synapsins: from synapse to network hyperexcitability and
and presynaptic plasticity in postnatal mouse brain. Neuro- epilepsy. Semin Cell Dev Biol. 22:408–415.
science. 158:231–241. Feliciano P, Matos H, Andrade R, Bykhovskaia M. 2017. Synapsin
Bornschein G, Schmidt H. 2018. Synaptotagmin Ca2+ sensors II regulation of GABAergic synaptic transmission is depen-
and their spatial coupling to presynaptic Cav channels in dent on interneuron subtype. J Neurosci Off J Soc Neurosci.
central cortical synapses. Front Mol Neurosci. 11:494. 37:1757–1771.
Cesca F, Baldelli P, Valtorta F, Benfenati F. 2010. The synapsins: Gan J, Weng S-M, Pernía-Andrade AJ, Csicsvari J, Jonas P. 2017.
key actors of synapse function and plasticity. Prog Neurobiol. Phase-locked inhibition, but not excitation, underlies hip-
91:313–348. pocampal ripple oscillations in awake mice in vivo. Neuron.
Chang S, Trimbuch T, Rosenmund C. 2018. Synaptotagmin-1 93:308–314.
drives synchronous Ca2+-triggered fusion by C2B-domain- Garcia CC. 2004. Identification of a mutation in synapsin I, a
mediated synaptic-vesicle-membrane attachment. Nat Neu- synaptic vesicle protein, in a family with epilepsy. J Med Genet.
rosci. 21:33–40. 41:183–186.
Chen C, Jonas P. 2017. Synaptotagmins: That’s why so many. Geppert M, Goda Y, Hammer RE, Li C, Rosahl TW, Stevens CF,
Neuron. 94:694–696. Südhof TC. 1994. Synaptotagmin I: a major Ca2+ sensor for
Chen C, Satterfield R, Young SM, Jonas P. 2017. Triple function transmitter release at a central synapse. Cell. 79:717–727.
of Synaptotagmin 7 ensures efficiency of high-frequency Giannandrea M, Guarnieri FC, Gehring NH, Monzani E, Ben-
transmission at central GABAergic synapses. Cell Rep. fenati F, Kulozik AE, Valtorta F. 2013. Nonsense-mediated
21:2082–2089. mRNA decay and loss-of-function of the protein underlie the
Chiappalone M, Casagrande S, Tedesco M, Valtorta F, Baldelli P, X-linked epilepsy associated with the W356× mutation in
Martinoia S, Benfenati F. 2009. Opposite changes in gluta- synapsin I. PloS One. 8:e67724.
matergic and GABAergic transmission underlie the diffuse Guarnieri FC, Pozzi D, Raimondi A, Fesce R, Valente MM, Delvec-
hyperexcitability of synapsin I-deficient cortical networks. chio VS, Van Esch H, Matteoli M, Benfenati F, D’Adamo P et al.
Cereb Cortex. 19:1422–1439. 2017. A novel SYN1 missense mutation in non-syndromic X-
Corradi A, Fadda M, Piton A, Patry L, Marte A, Rossi P, Cadieux- linked intellectual disability affects synaptic vesicle life cycle,
Dion M, Gauthier J, Lapointe L, Mottron L et al. 2014. SYN2 clustering and mobility. Hum Mol Genet. 26:4699–4714.
is an autism predisposing gene: loss-of-function mutations Hagler DJ, Goda Y. 2001. Properties of synchronous and asyn-
alter synaptic vesicle cycling and axon outgrowth. Hum Mol chronous release during pulse train depression in cultured
Genet. 23:90–103. hippocampal neurons. J Neurophysiol. 85:2324–2334.
14 Cerebral Cortex, 2019, Vol. 00, No. 00

Hefft S, Jonas P. 2005. Asynchronous GABA release gener- bathing seizures: characterization of a distinct epileptic syn-
ates long-lasting inhibition at a hippocampal interneuron- drome. Epilepsia. 56:1098–1108.
principal neuron synapse. Nat Neurosci. 8:1319–1328. Nishiki T, Augustine GJ. 2004. Synaptotagmin I synchronizes
Hu H, Gan J, Jonas P. 2014. Fast-spiking, parvalbumin+ GABAergic transmitter release in mouse hippocampal neurons. J Neu-

Downloaded from https://academic.oup.com/cercor/advance-article-abstract/doi/10.1093/cercor/bhz174/5546054 by Airlangga University user on 07 November 2019


interneurons: from cellular design to microcircuit function. rosci Off J Soc Neurosci. 24:6127–6132.
Science. 345:1255263–1255263. Otsu Y. 2004. Competition between phasic and asynchronous
Iremonger KJ, Bains JS. 2016. Asynchronous presynaptic gluta- release for recovered synaptic vesicles at developing hip-
mate release enhances neuronal excitability during the post- pocampal Autaptic synapses. J Neurosci. 24:420–433.
spike refractory period. J Physiol. 594:1005–1015. Peron A, Baratang NV, Canevini MP, Campeau PM, Vignoli
Isaacson JS, Walmsley B. 1995. Counting quanta: direct measure- A. 2018. Hot water epilepsy and SYN1 variants. Epilepsia.
ments of transmitter release at a central synapse. Neuron. 59:2162–2163.
15:875–884. Pyott SJ, Rosenmund C. 2002. The effects of temperature on
Jiang M, Zhu J, Liu Y, Yang M, Tian C, Jiang S, Wang Y, Guo H, vesicular supply and release in autaptic cultures of rat and
Wang K, Shu Y. 2012. Enhancement of asynchronous release mouse hippocampal neurons. J Physiol. 539:523–535.
from fast-spiking interneuron in human and rat epileptic Raingo J, Khvotchev M, Liu P, Darios F, Li YC, Ramirez DMO,
neocortex. PLoS Biol. 10:e1001324. Adachi M, Lemieux P, Toth K, Davletov B et al. 2012. VAMP4
Kaeser PS, Regehr WG. 2014. Molecular mechanisms for syn- directs synaptic vesicles to a pool that selectively main-
chronous, asynchronous, and spontaneous neurotransmitter tains asynchronous neurotransmission. Nat Neurosci. 15:
release. Ann Rev Physiol. 76:333–363. 738–745.
Labrakakis C, Lorenzo L-E, Bories C, Ribeiro-da-Silva A, De Rosahl TW, Geppert M, Spillane D, Herz J, Hammer RE, Malenka
Koninck Y. 2009. Inhibitory coupling between inhibitory RC, Südhof TC. 1993. Short-term synaptic plasticity is altered
interneurons in the spinal cord dorsal horn. Mol Pain. 5:24. in mice lacking synapsin. Cell. 75:661–670.
Lignani G, Raimondi A, Ferrea E, Rocchi A, Paonessa F, Cesca Rosahl TW, Spillane D, Missler M, Herz J, Selig DK, Wolff JR, Ham-
F, Orlando M, Tkatch T, Valtorta F, Cossette P et al. 2013. mer RE, Malenka RC, Südhof TC. 1995. Essential functions
Epileptogenic Q555X SYN1 mutant triggers imbalances in of synapsins I and II in synaptic vesicle regulation. Nature.
release dynamics and short-term plasticity. Hum Mol Genet. 375:488–493.
22:2186–2199. Ruiz R, Casañas JJ, Torres-Benito L, Cano R, Tabares L. 2010.
Liguz-Lecznar M, Urban-Ciecko J, Kossut M. 2016. Somatostatin Altered intracellular Ca2+ homeostasis in nerve terminals
and Somatostatin-containing neurons in shaping neuronal of severe spinal muscular atrophy mice. J Neurosci Off J Soc
activity and plasticity. Front Neural Circuits. 10:48. Neurosci. 30:849–857.
Lu T, Trussell LO. 2000. Inhibitory transmission mediated by Savanthrapadian S, Meyer T, Elgueta C, Booker SA, Vida I, Bartos
asynchronous transmitter release. Neuron. 26:683–694. M. 2014. Synaptic properties of SOM- and CCK-expressing
Manseau F, Marinelli S, Méndez P, Schwaller B, Prince DA, Hugue- cells in dentate Gyrus interneuron networks. J Neurosci.
nard JR, Bacci A. 2010. Desynchronization of neocortical net- 34:8197–8209.
works by asynchronous release of GABA at autaptic and Schneggenburger R, Meyer AC, Neher E. 1999. Released fraction
synaptic contacts from fast-spiking interneurons. PLoS Biol. and Total size of a Pool of immediately available transmitter
8:e1000492. quanta at a calyx synapse. Neuron. 23:399–409.
Matos M, Bosson A, Riebe I, Reynell C, Vallée J, Laplante I, Panatier Song S-H, Augustine GJ. 2016. Synapsin isoforms regulating
A, Robitaille R, Lacaille J-C. 2018. Astrocytes detect and upreg- GABA release from hippocampal interneurons. J Neurosci Off
ulate transmission at inhibitory synapses of somatostatin J Soc Neurosci. 36:6742–6757.
interneurons onto pyramidal cells. Nat Commun. 9:4254. Taniguchi H, He M, Wu P, Kim S, Paik R, Sugino K, Kvitsani D,
Maximov A, Tang J, Yang X, Pang ZP, Südhof TC. 2009. Complexin Fu Y, Lu J, Lin Y et al. 2011. A resource of Cre driver lines
controls the force transfer from SNARE complexes to mem- for genetic targeting of GABAergic neurons in cerebral cortex.
branes in fusion. Science. 323:516–521. Neuron. 71:995–1013.
Medrihan L, Cesca F, Raimondi A, Lignani G, Baldelli P, Benfenati Tremblay R, Lee S, Rudy B. 2016. GABAergic interneurons in
F. 2013. Synapsin II desynchronizes neurotransmitter release the neocortex: from cellular properties to circuits. Neuron.
at inhibitory synapses by interacting with presynaptic cal- 91:260–292.
cium channels. Nat Commun. 4:1512. Valente P, Farisello P, Valtorta F, Baldelli P, Benfenati F. 2017.
Medrihan L, Ferrea E, Greco B, Baldelli P, Benfenati F. 2015. Impaired GABAB-mediated presynaptic inhibition increases
Asynchronous GABA release is a key determinant of tonic excitatory strength and alters short-term plasticity in
inhibition and controls neuronal excitability: a study in the synapsin knockout mice. Oncotarget. 8:90061–90076.
Synapsin II-/- mouse. Cereb Cortex. 1991, 25:3356–3368. Yang L, Wang B, Long C, Wu G, Zheng H. 2007. Increased
Michetti C, Caruso A, Pagani M, Sabbioni M, Medrihan L, David asynchronous release and aberrant calcium channel activa-
G, Galbusera A, Morini M, Gozzi A, Benfenati F et al. 2017. The tion in amyloid precursor protein deficient neuromuscular
knockout of Synapsin II in mice impairs social behavior and synapses. Neuroscience. 149:768–778.
functional connectivity generating an ASD-like phenotype. Yao J, Gaffaney JD, Kwon SE, Chapman ER. 2011. Doc2 is a Ca2+
Cereb Cortex. 27:5014–5023. sensor required for asynchronous neurotransmitter release.
Mitchell SJ, Silver RA. 2003. Shunting inhibition modulates neu- Cell. 147:666–677.
ronal gain during synaptic excitation. Neuron. 38:433–445. Yavorska I, Wehr M. 2016. Somatostatin-expressing inhibitory
Nguyen DK, Rouleau I, Sénéchal G, Ansaldo AI, Gravel M, Ben- interneurons in cortical circuits. Front Neural Circuits.
fenati F, Cossette P. 2015. X-linked focal epilepsy with reflex 10:76.

Potrebbero piacerti anche