Autonomous Function of Synaptotagmin 1 in Triggering Synchronous Release Independent of Asynchronous Release

Similar documents
THE SYNAPTIC VESICLE CYCLE

Synaptotagmin-7-Mediated Asynchronous Release Boosts High-Fidelity Synchronous Transmission at a Central Synapse

Supplementary Materials for VAMP4 directs synaptic vesicles to a pool that selectively maintains asynchronous neurotransmission

Supplementary Figure 1. Basic properties of compound EPSPs at

Competition between Phasic and Asynchronous Release for Recovered Synaptic Vesicles at Developing Hippocampal Autaptic Synapses

Intraterminal Ca 2+ concentration and asynchronous transmitter release at single GABAergic boutons in rat collicular cultures

Neuroscience 201A (2016) - Problems in Synaptic Physiology

1) Drop off in the Bi 150 box outside Baxter 331 or to the head TA (jcolas).

Chapter 3 subtitles Action potentials

DOI: /jphysiol The Physiological Society Rapid Report

Synaptotagmin-1, -2, and -9: Ca 2+ Sensors for Fast Release that Specify Distinct Presynaptic Properties in Subsets of Neurons

Supporting Information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION. Supplementary Figure 1

Properties of Synchronous and Asynchronous Release During Pulse Train Depression in Cultured Hippocampal Neurons

Supplementary Figure 1. SybII and Ceb are sorted to distinct vesicle populations in astrocytes. Nature Neuroscience: doi: /nn.

BIPN 140 Problem Set 6

BIPN 140 Problem Set 6

MCB MIDTERM EXAM #1 MONDAY MARCH 3, 2008 ANSWER KEY

Synaptotagmin-1- and Synaptotagmin-7-Dependent Fusion Mechanisms Target Synaptic Vesicles to Kinetically Distinct Endocytic Pathways

Supporting Online Material for

Chapter 3 Neurotransmitter release

Readily releasable pool size changes associated with long term depression

Presynaptic NMDA receptor control of spontaneous and evoked activity By: Sally Si Ying Li Supervisor: Jesper Sjöström

Abstract. 1 Introduction

Impaired GABAergic inhibition in the visual cortex of brain-derived neurotrophic factor heterozygous knockout mice

Supplementary Information

Cellular Neurobiology / BIPN 140

Modelling Vesicular Release at Hippocampal Synapses

Modeling Depolarization Induced Suppression of Inhibition in Pyramidal Neurons

Decreased Frequency But Not Amplitude of Quantal Synaptic Responses Associated with Expression of Corticostriatal Long-Term Depression

Quantal Analysis Problems

Alterations in Synaptic Strength Preceding Axon Withdrawal

Activity Dependent Changes At the Developing Neuromuscular Junction

The effects of temperature on vesicular supply and release in autaptic cultures of rat and mouse hippocampal neurons

At the synapse, rapid release of neurotransmitter occurs when

Ivy/Neurogliaform Interneurons Coordinate Activity in the Neurogenic Niche

Part 11: Mechanisms of Learning

The control of spiking by synaptic input in striatal and pallidal neurons

Desynchronization of Neocortical Networks by Asynchronous Release of GABA at Autaptic and Synaptic Contacts from Fast-Spiking Interneurons

Synaptic communication

1) Drop off in the Bi 150 box outside Baxter 331 or to the head TA (jcolas).

Astrocyte signaling controls spike timing-dependent depression at neocortical synapses

Presynaptic control of e$cacy of GABAergic synapses in the hippocampus

Synaptic Transmission: Ionic and Metabotropic

When cells are already maximally potentiated LTP is occluded.

Cd k5 Ph o s p h o r y l at i o n o f Mu n c18-1 En h a n c e s Sy n a p t i c

File name: Supplementary Information Description: Supplementary Figures, Supplementary Table and Supplementary References

The molecular analysis of long-term plasticity in the mammalian

Sample Lab Report 1 from 1. Measuring and Manipulating Passive Membrane Properties

Synaptic Integration

Supplementary figure 1: LII/III GIN-cells show morphological characteristics of MC

The mammalian cochlea possesses two classes of afferent neurons and two classes of efferent neurons.

Autism-Associated Neuroligin-3 Mutations Commonly Disrupt Tonic Endocannabinoid Signaling

Problem Set 3 - Answers. -70mV TBOA

SYNAPTIC COMMUNICATION

Three distinct kinetic groupings of the synaptotagmin family: Candidate sensors for rapid and delayed exocytosis

Supplementary Figure 1) GABAergic enhancement by leptin hyperpolarizes POMC neurons A) Representative recording samples showing the membrane

Chapter 6 subtitles postsynaptic integration

Synaptic plasticityhippocampus. Neur 8790 Topics in Neuroscience: Neuroplasticity. Outline. Synaptic plasticity hypothesis

Primary Afferent Activation of Thermosensitive TRPV1 Triggers Asynchronous Glutamate Release at Central Neurons

Enhancement of synaptic transmission by cyclic AMP modulation of presynaptic I h channels. Vahri Beaumont and Robert S. Zucker

Synaptic Communication. Steven McLoon Department of Neuroscience University of Minnesota

Chapter 45: Synapses Transmission of Nerve Impulses Between Neurons. Chad Smurthwaite & Jordan Shellmire

Nature Neuroscience: doi: /nn.2275

Cholinergic Activation of M2 Receptors Leads to Context- Dependent Modulation of Feedforward Inhibition in the Visual Thalamus

LETTERS. Allosteric modulation of the presynaptic Ca 21 sensor for vesicle fusion. Xuelin Lou 1, Volker Scheuss 1 & Ralf Schneggenburger 1

A General Model of Synaptic Transmission and Short-Term Plasticity

QUIZ/TEST REVIEW NOTES SECTION 7 NEUROPHYSIOLOGY [THE SYNAPSE AND PHARMACOLOGY]

The Influence of Multivesicular Release and Postsynaptic Receptor Saturation on Transmission at Granule Cell to Purkinje Cell Synapses

Neurons! John A. White Dept. of Bioengineering

GABA B Receptor-Mediated Presynaptic Inhibition Has History-Dependent Effects on Synaptic Transmission during Physiologically Relevant Spike Trains

Charlie Taylor, PhD CpTaylor Consulting Chelsea, MI, USA

Effects of adrenaline on nerve terminals in the superior cervical ganglion of the rabbit

Synaptic Plasticity and Memory

Ube3a is required for experience-dependent maturation of the neocortex

Dep. Control Time (min)

DOMINIQUE DEBANNE*, BEAT H. GÄHWILER, AND SCOTT M. THOMPSON MATERIALS AND METHODS

Synaptic Plasticity and the NMDA Receptor

Portions from Chapter 6 CHAPTER 7. The Nervous System: Neurons and Synapses. Chapter 7 Outline. and Supporting Cells

The Journal of Neuroscience, December 15, 2002, 22(24):

Concept 48.1 Neuron organization and structure reflect function in information transfer

Lecture 22: A little Neurobiology

Supplementary Information

How Nicotinic Signaling Shapes Neural Networks

How Synapses Integrate Information and Change

Synaptic transmission

STRUCTURAL ELEMENTS OF THE NERVOUS SYSTEM

Ultrastructural Contributions to Desensitization at the Cerebellar Mossy Fiber to Granule Cell Synapse

Chapter 5 subtitles GABAergic synaptic transmission

Anatomy Review. Graphics are used with permission of: Pearson Education Inc., publishing as Benjamin Cummings (

How Synapses Integrate Information and Change

BIONB/BME/ECE 4910 Neuronal Simulation Assignments 1, Spring 2013

Spatial Distribution of Calcium Entry Evoked by Single Action Potentials within the Presynaptic Active Zone

Bidirectional NMDA receptor plasticity controls CA3 output and heterosynaptic metaplasticity

Summary of Calcium Regulation inside the Cell

Synaptic Transmission

Differing mechanisms of exocytosis for large dense core vesicles in chromaffin cells and small synaptic vesicles in dopamine neurons.

Excitatory synaptic transmission in the central nervous system. Mechanisms underlying kainate receptor-mediated disinhibition in the hippocampus

Transcription:

Neuron, Vol. 48, 547 554, November 23, 2005, Copyright ª2005 by Elsevier Inc. DOI 10.1016/j.neuron.2005.09.006 Autonomous Function of Synaptotagmin 1 in Triggering Synchronous Release Independent of Asynchronous Release Report Anton Maximov and Thomas C. Südhof* Center for Basic Neuroscience Department of Molecular Genetics and Howard Hughes Medical Institute The University of Texas Southwestern Medical Center Dallas, Texas 75390 Summary Ca 2+ triggers neurotransmitter release in at least two principal modes, synchronous and asynchronous release. Synaptotagmin 1 functions as a Ca 2+ sensor for synchronous release, but its role in asynchronous release remains unclear. We now show that in cultured cortical neurons stimulated at low frequency (%0.1 Hz), deletion of synaptotagmin 1 blocks synchronous GABA and glutamate release without significantly increasing asynchronous release. At higher stimulation frequencies (R4 Hz), deletion of synaptotagmin 1 also alters only synchronous, not asynchronous, release during the stimulus train, but dramatically enhances delayed asynchronous release following the stimulus train. Thus synaptotagmin 1 functions as an autonomous Ca 2+ sensor independent of asynchronous release during isolated action potentials and action potential trains, but restricts asynchronous release induced by residual Ca 2+ after action potential trains. We propose that synaptotagmin 1 occupies release slots at the active zone, possibly in a Ca 2+ - independent complex with SNARE proteins that are freed when action potential-induced Ca 2+ influx activates synaptotagmin 1. Introduction Neurons exhibit two modes of Ca 2+ -triggered neurotransmitter release. When action potentials fire at low frequency, most release is synchronous (also called phasic ). When action potentials fire at high frequency, release becomes largely asynchronous (Barrett and Stevens, 1972; Goda and Stevens, 1994; Cummings et al., 1996; Atluri and Regehr, 1998; Lu and Trussell, 2000; Hagler and Goda, 2001; Otsu et al., 2004; reviewed in Atwood and Karunanithi, 2002). Synchronous release is triggered by brief localized Ca 2+ waves induced by action potentials (Llinas et al., 1995), whereas asynchronous release is triggered by increased bulk residual Ca 2+ (Atluri and Regehr, 1998; Jensen et al., 2000; Kirischuk and Grantyn, 2003). Asynchronous release induced by high-frequency stimulation consists of two types: (1) release during the stimulus train when synchronous release still operates but is outcompeted by asynchronous release, and (2) release after the stimulus train ( delayed release ) when residual Ca 2+ remains sufficiently high to trigger synaptic vesicle exocytosis. *Correspondence: Thomas.Sudhof@UTSouthwestern.edu In addition to evoked synchronous and asynchronous release, neurons exhibit spontaneous neurotransmitter release independent of action potentials that will not be examined here. Synchronous and asynchronous release are derived from the same pool of readily releasable vesicles (Lu and Trussell, 2000; Otsu et al., 2004). During highfrequency stimulation, release switches from synchronous to asynchronous when the readily releasable pool becomes depleted, and the accumulating residual Ca 2+ triggers asynchronous exocytosis of vesicles as they refill the pool. Under these conditions, action potentials still induce waves of Ca 2+ influx. Nevertheless, asynchronous release outcompetes synchronous release for recovered quanta, probably because asynchronous release is triggered by a slow, high-affinity Ca 2+ sensor that operates in the intervals between the action potentials. In contrast, delayed release is induced by residual Ca 2+ after the stimulus train ends, when no additional waves of Ca 2+ influx occur. At least in some synapses, asynchronous release may be important to sustain synaptic transmission during high-frequency stimulus trains (Lu and Trussell, 2000), but the overall biological significance of asynchronous release remains unclear. In excitatory synapses of the hippocampus, synchronous release is triggered by Ca 2+ binding to synaptotagmin 1 (Geppert et al., 1994; Fernández-Chacón et al., 2001). Synaptotagmin 1 probably functions similarly in Drosophila as synaptotagmin 1 is highly conserved (Perin et al., 1991) and mutations of synaptotagmin 1 in Drosophila have a similar phenotype (Di Antonio et al., 1993; Littleton et al., 1993), although this conclusion is not universally accepted (Robinson et al., 2002). A major question is how synchronous release triggered by Ca 2+ binding to synaptotagmin 1 relates to asynchronous release. One possibility is that independent Ca 2+ sensors control synchronous and asynchronous release, as suggested in the analysis of synaptotagmin 1 mutant mice (Geppert et al., 1994) and of release induced by highfrequency stimulation (Hagler and Goda, 2001; Otsu et al., 2004). However, in hippocampal autapses the total amount of release triggered by an action potential, when measured over a full second, is not dramatically different in the presence and absence of synaptotagmin 1, and only the time course of release differs (Shin et al., 2003; Nishiki and Augustine, 2004). Similarly, release triggered by low Ca 2+ concentrations in the absence of action potentials is enhanced in synaptotagmin 1 mutants in Drosophila (Yoshihara and Littleton, 2002). These findings suggested that deletion of synaptotagmin 1 only changes the relative amount of synchronous versus asynchronous release, but not the total amount of release. Based on this conclusion, a second hypothesis of synaptotagmin 1 function was proposed, namely, that synaptotagmin 1 synchronizes release that would normally happen anyway (Yoshihara and Littleton, 2002; Nishiki and Augustine, 2004; reviewed in Tokuoka and Goda, 2003). To date, analyses of synaptotagmin 1 function were restricted to autapses formed by cultured hippocampal

Neuron 548 neurons, neuromuscular junctions of Drosophila, and chromaffin cells (e.g., see Geppert et al., 1994; Yoshihara and Littleton, 2002; Shin et al., 2003; Nishiki and Augustine, 2004; Voets et al., 2001; Sorensen et al., 2003). Surprisingly, the role of synaptotagmin 1 has not been studied in detail in an interneuronal synapse, the primary site of action of synaptotagmin 1. Furthermore, the role of synaptotagmin 1 at inhibitory synapses is unknown, and thus the generality of synaptotagmin 1 as a Ca 2+ sensor for synchronous release remains uncertain. Finally, the importance of synaptotagmin 1 for release induced by high-frequency stimulus trains that physiologically occur in many synapses has not been investigated. In investigating these questions, we now find that deletion of synaptotagmin 1 severely impaired release at both inhibitory and excitatory interneuronal synapses. Synaptotagmin 1 deficiency had no major effect on asynchronous release induced by isolated action potentials and did not significantly alter asynchronous release during high-frequency stimulus trains. However, deletion of synaptotagmin 1 dramatically increased delayed release. We propose that synchronous and asynchronous release are independent of each other when release is triggered by action potential-induced Ca 2+ influx and when synaptotagmin 1 as the Ca 2+ sensor for synchronous release competes with unknown Ca 2+ sensors for asynchronous release. However, when release is triggered by residual Ca 2+ without action potential-induced Ca 2+ influx and without Ca 2+ binding to synaptotagmin 1, synaptotagmin 1 restricts asynchronous release. Results Excitatory Synaptic Transmission in Cortical Neurons from Wild-Type and Synaptotagmin 1 KO Mice We analyzed synaptic transmission in mixed cortical neurons cultured from littermate wild-type and synaptotagmin 1 knockout (KO) mice (Figure S1 in the Supplemental Data available with this article online). To monitor synaptic transmission between neurons, we placed an extracellular electrode near the cell body of a neuron, induced action potentials by current injections (900 ma for 1 ms), and measured postsynaptic currents in a neighboring neuron with whole-cell recordings. Wild-type neurons exhibited typically large amplitudes of AMPA-type excitatory postsynaptic currents (EPSCs) in response to action potentials, whereas synaptotagmin 1 KO neurons displayed w10-fold smaller EPSC amplitudes (Figure 1A). To compare the time course of release between wild-type and synaptotagmin 1-deficient neurons, we plotted the time-dependence of the number of quanta released, measured as the charge transfer during 10 ms intervals divided by the average charge of spontaneous release events (Figure 1B). This plot revealed that release at <300 ms after an action potential was almost absent in synaptotagmin 1-deficient synapses, while release at >300 ms, the time period at which asynchronous release predominates after a single action potential (Goda and Stevens, 1994), did not change significantly. Synaptotagmin 1 Triggers Synchronous Release in Inhibitory Synapses To test whether synaptotagmin 1 also functions in inhibitory synapses, we monitored evoked inhibitory postsynaptic currents (IPSCs) in the presence of the AMPA receptor blocker CNQX (20 mm) and the NMDA receptor blocker AP-5 (50 mm). We observed robust IPSCs in wild-type neurons that were blocked with picrotoxin (Figure 1C). In synaptotagmin 1-deficient neurons, inhibitory synaptic responses were severely impaired. Both the IPSC amplitudes and the total synaptic charge transfer over 1.5 s after an action potential were decreased w10-fold in mutant synapses (Figure 1D). Again, we observed no significant increase in asynchronous release in synaptotagmin 1-deficient synapses. IPSCs in synaptotagmin 1-deficient neurons were not only smaller, but also displayed an w30-fold increase in rise time and a large variation in onset (Figure 1C, insets and Figure 1F). We estimated the speed of release in 10 mm Ca 2+ by integrating the total charge transfer during single IPSCs in wild-type and synaptotagmin 1-deficient neurons over 1.5 s (Figure S2). The release time course thus determined can be fitted to a doubleexponential function in wild-type neurons (t1 = 111 6 8 ms; t2 = 406 6 12 ms; n = 11), corresponding to the fast and slow components of release, respectively. In synaptotagmin 1-deficient neurons, however, singleexponential fits were sufficient, yielding a time constant similar to that of the slow release component (t = 466 6 17 ms; n = 19). Together, the decreased rise times, variable latency, and slow time course of the residual release in synaptotagmin 1-deficient synapses suggest that this residual release is predominantly asynchronous. A major concern in the analysis of a KO phenotype is that the changes observed do not reflect the acute function of a protein, but its developmental history. To address this concern, we examined IPSCs in neurons that were infected with a lentivirus-expressing synaptotagmin 1. All aspects of the KO phenotype were rescued: the kinetics of release (Figure 1C and Figure S2), the IPSC amplitudes (Figure S3), and the synchronization and rise times of responses (Figure 1F). Thus, synaptotagmin 1 performs an essential function in Ca 2+ - triggered release in inhibitory synapses similar to excitatory synapses, and this function is not developmentally determined but acutely operates in the nerve terminal during release. Asynchronous Release in Synaptotagmin 1-Deficient Synapses Is Sensitive to EGTA We next examined how EGTA-AM, an EGTA ester that is hydrolyzed to the slow Ca 2+ buffer EGTA in cells, affects release. Wild-type neurons preincubated with EGTA-AM (0.1 mm for 3 5 min) exhibited a moderate decrease in the IPSC amplitude and an acceleration of the IPSC decay (Figures 1D and 1H). These changes appeared to be presynaptic because we observed no significant alterations in the kinetics of spontaneous release events by increasing the Ca 2+ concentration or by addition of EGTA- AM (Figure S4; note that the frequency of spontaneous mipscs, measured in 10 mm extracellular Ca 2+, was increased from 3.4 6 0.4 Hz [n = 6] in wild-type to 11 6 0.6 Hz [n = 14] in mutant neurons; this increase will be examined in a later study). In contrast to its moderate

Role of Synaptotagmin 1 in Asynchronous Release 549 Figure 1. Role of Synaptotagmin 1 in Excitatory and Inhibitory Synaptic Transmission during 0.1 Hz Stimulation (A) Representative AMPA-type EPSCs evoked by extracellular stimulation in 10 mm extracellular Ca 2+ in cultured cortical neurons from wild-type (WT) and synaptotagmin 1 KO mice (Syt 1 KO) (14 days in vitro; arrow indicates time of stimulation). (B) Quantal release in wild-type and synaptotagmin 1-deficient neurons. The number of quanta released per action potential was estimated by averaging the charge transfer during 50 individual responses collected at 0.1 Hz. Charges were integrated over 10 ms bins and divided by the average charge transfer of spontaneous EPSCs (100 fc). Insert illustrates the average EPSC amplitudes in wild-type and synaptotagmin 1-deficient neurons (n = 3; *p < 0.05). (C) Representative IPSCs recorded in wildtype neurons, naive synaptotagmin 1-deficient neurons, and synaptotagmin 1-deficient neurons infected with a lentivirus that expresses full-length synaptotagmin 1. Recordings were performed in 10 mm extracellular Ca 2+ in the presence of AP-5 and CNQX. Note that the wild-type IPSCs can be completely blocked by picrotoxin (PTX). The inserts below the traces show expansions of the initial phase of individual IPSCs to illustrate the variability of onset in responses recorded from synaptotagmin 1-deficient neurons. (D G) Quantitation of the properties of EPSC and IPSC amplitudes (top panels) and synaptic charge transfer (bottom panels). (D) depicts mean IPSC amplitudes and charge transfers in wild-type (WT) or synaptotagmin 1 KO synapses in 2 mm or 10 mm Ca 2+,asindicated, in the absence or presence of 0.1 mm EGTA-AM. (E) displays the ratios of IPSC amplitudes and charges evoked in EGTA-AMtreated versus untreated synapses, and (F) illustrates the standard deviation of time-topeak (top) or rise times (bottom) of IPSCs in wild-type and KO synapses and KO synapses from neurons after lentiviral rescue (all in 10 mm Ca 2+ ). (G) shows EPSC amplitudes and charges for KO synapses in the absence and presence of EGTA. Error bars indicate mean 6 SEM; number of neurons analyzed: (D and E): WT, n = 21; WT/EGTA, n = 6; KO 2mMCa 2+, n = 6; KO/EGTA 2 mm Ca 2+,n=3; KO 10 mm Ca 2+, n = 28; KO/EGTA 10 mm Ca 2+, n = 9; (F [top]): WT, n = 32; KO, n = 14; KO rescued, n = 3; (F [bottom]): WT, n = 7; KO, n = 17; KO rescued, n = 6; (G): KO, n = 9; KO/EGTA, n = 3. Asterisks indicate statistically significant differences for KO versus wild-type: *p < 0.05; **p < 0.01; ***p < 0.001. (H) Representative IPSCs recorded from the same wild-type or synaptotagmin 1-deficient neurons in 10 mm extracellular Ca 2+ before and 2 min after addition of 0.1 mm EGTA-AM. (I) Representative EPSCs recorded in 10 mm extracellular Ca 2+ from synaptotagmin 1-deficient neurons before and after addition of 0.1 mm EGTA- AM. All data shown represent means 6 SEM. effects in wild-type neurons, EGTA-AM nearly abolished the residual release that remains in synaptotagmin 1-deficient neurons (Figures 1D and 1H), as evidenced by the low ratio of EGTA-resistant release to total release (Figure 1E). We observed a similarly dramatic suppression by EGTA-AM on EPSCs in synaptotagmin 1-deficient neurons (Figures 1G and 1I). Since previous studies showed that asynchronous release is much more

Neuron 550 sensitive to EGTA-AM than synchronous release (Otsu et al., 2004), and since we used relatively low concentrations of EGTA-AM (0.1 mm) but a high extracellular Ca 2+ concentration (10 mm), these results indicate that all of the residual release in synaptotagmin 1-deficient neurons is probably asynchronous. Figure 2. Role of Synaptotagmin 1 during Repetitive Stimulation (A) Representative traces of IPSCs of wild-type neurons stimulated by 50 action potentials applied at 10 Hz. Responses were recorded in the presence of 2 mm extracellular Ca 2+ (top), 10 mm Ca 2+ (middle), or 10 mm Ca 2+ after addition of 0.1 mm EGTA-AM (bottom). Arrows indicate the end of the stimulus train. Calibration bars apply to all traces. (B) Time-to-peak (normalized for the initial response) of individual IPSCs during 20 Hz stimulus trains plotted as a function of stimulus number (n = 3). (C) Representative IPSCs and EPSCs from synaptotagmin 1-deficient neurons stimulated by 50 action potentials at 10 Hz in 10 mm Ca 2+ before or after addition of 0.1 mm EGTA-AM. Arrows indicate the end of the stimulus train. Calibration bar on the right applies to all traces. (D and E) Average synaptic charge transfer per action potential calculated for inhibitory responses from wild-type (D) and synaptotagmin 1-deficient neurons (E) during a 10 Hz stimulus train. For each time interval between action potentials, the total charge transfer including asynchronous release was calculated (see inset in [D]). Recordings were performed in 10 mm Ca 2+ without (control) or with 0.1 mm EGTA-AM ([D]: control, Role of Synaptotagmin 1 during Repetitive Stimulation To test the role of synaptotagmin 1 in Ca 2+ -triggered release evoked by high-frequency stimulus trains, we examined IPSCs induced by 50 action potentials applied at 10 Hz in 2 mm or 10 mm extracellular Ca 2+ (Figure 2A). In wild-type neurons, the first synaptic response was synchronous. Starting with the second action potential, however, synchronous IPSCs were superimposed on a background current that is likely due to asynchronous release (Otsu et al., 2004). Synchronous IPSCs riding on top of the asynchronous component were discernable throughout the stimulus train in 2 mm Ca 2+, but were occluded in 10 mm Ca 2+ (Figure 2A). EGTA-AM nearly eliminated asynchronous release during the stimulus train and resynchronized synaptic responses (Figure 2A and Figure S5). Similar observations were made at other stimulation frequencies (data not shown). The synchronization of release by EGTA was visualized by plotting the time-to-peak of IPSCs as a function of stimulus number (Figure 2B). During stimulus trains, the shapes of IPSCs become more and more irregular in the absence of EGTA, but remain uniform in the presence of EGTA (Figure S5). Although EGTA-AM partly lowers synchronous release in addition to suppressing asynchronous release (Figures 1C and 1D), we detected no facilitation of responses during the stimulus train by EGTA-AM (Figure 2A), presumably because residual Ca 2+ is required for facilitation (Zucker and Regehr, 2002). After the stimulus train, we observed a tail current that corresponds to delayed asynchronous release (Atluri and Regehr, 1998). Similar to asynchronous release during the stimulus train, this tail current was abolished by EGTA-AM similar to asynchronous release during the stimulus train (Figure 2A and Figures S6 and S7). In synaptotagmin 1-deficient neurons, initial responses at the beginning of a stimulus train were notably decreased compared to wild-type neurons, whereas the remaining synaptic currents were similar (Figure 2C). The large size of initial responses in wild-type synapses versus their slow buildup in synaptotagmin 1-deficient synapses is illustrated by plotting the synaptic charge transfer as a function of the stimulus number (Figures 2D and 2E). The loss of the initial release component, but not of steady-state release, during the stimulus train agrees well with the selective loss of synchronous, but not asynchronous, release in synaptotagmin 1-deficient neurons. EGTA-AM suppressed asynchronous release during the stimulus train and delayed release in the synaptotagmin 1-deficient neurons similarly to its effects on wild-type neurons (Figure 2C). However, unlike the case in wild-type neurons, EGTA-AM did not synchronize n = 5; EGTA-AM, n = 6; [E]: control, n = 11; EGTA-AM, n = 11). The smooth lines represent the difference between the responses in the absence and presence of EGTA. All data shown represent means 6 SEM.

Role of Synaptotagmin 1 in Asynchronous Release 551 Figure 3. Deletion of Synaptotagmin 1 Causes a Selective Increase in Delayed Release IPSCs were triggered by trains of 2 30 action potentials applied at 4 Hz (A) or 10 Hz (B) in 2 mm Ca 2+. The current traces illustrate typical IPSCs recorded from wild-type (red) and synaptotagmin 1-deficient neurons (blue) triggered by trains of eight action potentials. The scale bars apply to all traces. The arrows indicate the tail currents (delayed release) observed at the end of stimulus trains. Average charges were calculated separately for release observed during repetitive stimulation (top panels), delayed release (middle panels), and total release (bottom panels) and were plotted as a function of the number of action potentials in a stimulus train. A: wild-type, n = 8; synaptotagmin 1 KO, n = 8; B: wildtype, n = 11; synaptotagmin 1 KO, n = 7, synaptotagmin 1 KO with EGTA-AM, n = 3. For experiments analyzing delayed release in 10 mm Ca 2+ in response to stimulation frequencies of 5 50 Hz, see Figure S6. All data shown represent means 6 SEM. synaptic responses in synaptotagmin 1-deficient neurons, presumably because the synchronous component of release is missing. Interestingly, if we subtract the EGTA-AM-resistant release from the control release in wild-type neurons (Figure 2D), we observe a time course for the EGTA-sensitive release component that except for the initial responses closely resembles the time course of release in synaptotagmin 1-deficient neurons (Figure 2E). Selective Augmentation of Delayed Release by Deletion of Synaptotagmin 1 We next systematically studied the amount of synaptic charge transfer during and after stimulus trains of different lengths and frequencies (Figure 3 and Figures S6 and S7). When we applied 2 30 action potentials at 4 Hz or 10 Hz in 2 mm extracellular Ca 2+, the amount of release during the stimulus trains was identical between wild-type and synaptotagmin 1-deficient neurons. Surprisingly, however, delayed release after the stimulus trains was increased in synaptotagmin 1-deficient neurons. After only two action potentials, delayed release was augmented; after w10 action potentials, delayed release reached a plateau that was 2- to 5-fold higher in mutant than in wild-type synapses (Figure 3 and Figure S6). The increase in delayed release with no change in stimulus train-induced release was also observed when we enhanced the release probability by raising the extracellular Ca 2+ concentration, was found at all tested stimulation frequencies (5 50 Hz; Figure S6), and was associated with an increase in decay constants (Figure S7). This observation is not simply due to nonspecific damage of the mutant synapses by highfrequency stimulation, because EGTA-AM almost completely blocked release during and after the stimulus train in wild-type and mutant neurons (Figure 3). Release Induced by Directly Introducing Ca 2+ into Nerve Terminals Using Ionomycin The increase of delayed release in synaptotagmin 1- deficient synapses indicates that in these synapses, residual Ca 2+ remaining after a stimulus train (but not during the stimulus train) may be more potent in triggering asynchronous release. However, it is also possible that a difference in the clearance of residual Ca 2+ or an unexpected role for synaptotagmin 1 in inhibiting vesicle recycling may cause the increase of delayed release. Although the fact that only two action potentials at 4 Hz are sufficient to induce increased delayed release in mutant synapses argues against these alternative explanations,

Neuron 552 we tested this issue further by measuring the kinetics and size of IPSCs induced by addition of the Ca 2+ ionophore ionomycin (Figure 4). When ionomycin is added to the medium, it spontaneously incorporates into the neuronal plasma membrane, allowing Ca 2+ to directly enter nerve terminals. We found that synaptotagmin 1-deficient synapses responded to ionomycin almost twice as fast as wild-type neurons, whereas the total amount of release (which presumably corresponds to the readily releasable pool) was identical (Figure 4). This result supports the hypothesis that uniformly increased concentrations of bulk intracellular Ca 2+ trigger asynchronous release in synaptotagmin 1-deficient synapses at lower concentrations but in the same quantity in comparison with wild-type synapses. Discussion Figure 4. Effect of Ionomycin on Release in Wild-Type and Synaptotagmin 1-Deficient Synapses (A) Typical IPSCs triggered by application of 5 mm ionomycin in wildtype (red) and synaptotagmin 1-deficient neurons (blue) in 1 mm Ca 2+. Excitatory currents were blocked by AP5 and CNQX. The scale bars apply to both traces. (B) Normalized cumulative charge as a function of time to illustrate the kinetics of ionomycin-induced release in wild-type and synaptotagmin 1-deficient synapses. (C) Average delays from the point of application of ionomycin to maximum amplitude of inhibitory response in wild-type and synaptotagmin 1- deficient neurons (left); rise times from 10% of ionomycin-induced response to maximum response (middle), and total synaptic charge transfer for ionomycin-induced responses recorded from wild-type and synaptotagmin 1-deficient neurons (B and C, right: wild-type, n = 11; synaptotagmin 1 KO, n = 9). All data shown represent means 6 SEM. We employed electrophysiological recordings from interneuronal synapses in mixed cultures of cortical neurons to show that synaptotagmin 1 is essential for Ca 2+ triggering of synchronous release at both excitatory and inhibitory synapses, but is not essential for asynchronous release at these synapses. We then examined synchronous and asynchronous release in wild-type and synaptotagmin 1-deficient neurons under three conditions: (1) In response to isolated action potentials, (2) during action potential trains, and (3) during delayed release following action potential trains. Under the first condition, in response to isolated action potentials, release in wild-type neurons was almost entirely synchronous, but dramatically depressed in synaptotagmin 1-deficient neurons. The small amount of residual release in mutant synapses was primarily asynchronous, as shown by the increased rise times, variable onset, and EGTA sensitivity (Figure 1). Although we cannot rule out a small increase in asynchronous release in synaptotagmin 1-deficient neurons, any such increase must be minor because the total release in synaptotagmin 1-deficient neurons was so low. Thus, deletion of synaptotagmin 1 does not disinhibit asynchronous release induced by action potentials in cortical excitatory or inhibitory synapses. This is in contrast to previous observations in autapses where we (Shin et al., 2003) and others (Nishiki and Augustine, 2004) observed that the deletion of synaptotagmin 1 dramatically changes the time course of release, but has a much lesser moderate effect on the total amount of release integrated over 1 s. During action potential trains (the second condition), release in wild-type neurons was initially synchronous, but quickly became almost entirely asynchronous. Release in synaptotagmin 1-deficient neurons lacked the initial synchronous phase, but retained the dominant asynchronous phase. As a result, the total amount of release during stimulus trains, different from the release induced by isolated action potentials, was indistinguishable between synaptotagmin 1-deficient and wild-type neurons. Thus, when a steady train of action potentials pulses the nerve terminal with Ca 2+ but at the same time releases switches into a primarily asynchronous mode, synaptotagmin 1-deficient neurons respond remarkably similarly to wild-type neurons. Under the third condition, delayed release following action potential trains, deletion of synaptotagmin 1 dramatically increased asynchronous release independent of the stimulation frequency, number of action potentials, and extracellular Ca 2+ concentration. Thus, deletion of synaptotagmin 1 disinhibits asynchronous release under conditions in which an elevated level of residual Ca 2+ (accumulated during the stimulus train) triggers asynchronous release, but no additional pulses of peak Ca 2+ are added by action potentials. Based on these findings, we conclude that synaptotagmin 1 does not directly affect the action of the Ca 2+ sensor for asynchronous release in interneuronal synapses during action potential-induced release and

Role of Synaptotagmin 1 in Asynchronous Release 553 does not simply synchronize release. This conclusion is in agreement with previous findings that the Ca 2+ sensors for synchronous and asynchronous release operate independently (Lu and Trussell, 2000; Otsu et al., 2004) and that in chromaffin cells mutations of synaptotagmin 1 selectively impair the fast component of Ca 2+ -triggered exocytosis without changing the size of the slower components (Voets et al., 2001; Sorensen et al., 2003). The lack of a change in total release during the stimulus train in synaptotagmin 1-deficient synapses is probably not due to saturation of responses, since it was already observed after a few stimuli. However, synaptotagmin 1 does alter Ca 2+ triggering of asynchronous release because it dramatically increases delayed release. This observation suggests a mechanistic difference between asynchronous release during and after a stimulus train and could be explained by at least three hypotheses. First, synaptotagmin 1 normally suppresses Ca 2+ triggering of asynchronous release when the cytoplasmic bulk Ca 2+ concentration is elevated, but no action potential-induced waves of Ca 2+ influx occur. According to this hypothesis, synaptotagmin 1 only suppresses asynchronous release when it itself does not bind Ca 2+ after an action potential. Second, synaptotagmin 1 normally decreases residual Ca 2+ after a stimulus train, leading to an increase in residual Ca 2+ in mutant synapses and thus more delayed release. Third, synaptotagmin 1 normally inhibits vesicle endocytosis, recycling, and/or repriming; as a result, deletion of synaptotagmin 1 causes a relative increase in available vesicles after a stimulus train. We tested these mechanisms in stimulation experiments that applied a large range of action potential numbers (1 50), frequencies (4 50 Hz), and two different concentrations of extracellular Ca 2+. The results demonstrated that increased delayed release in mutant synapses is independent of the release probability or the number of action potentials, and they strongly argue against the second and third hypotheses. Moreover, we directly monitored release after addition of ionomycin, which uniformly increases the bulk Ca 2+ concentration in nerve terminals in a time-dependent manner, and found that mutant synapses responded much earlier during Ca 2+ influx (i.e., at lower Ca 2+ concentrations), but exhibited the same amount of total release (consistent with previous data showing that the size of the readily releasable pool is unaltered in synaptotagmin 1-deficient neurons; Geppert et al., 1994). Together, these findings indicate that synaptotagmin 1 impedes asynchronous release only when it itself does not become Ca 2+ bound, but it has no effect on asynchronous release when it becomes Ca 2+ bound even if Ca 2+ binding to synaptotagmin 1 does not actually trigger synchronous release under that particular condition. A plausible hypothesis to explain these surprising findings is based on the previous observation that synchronous and asynchronous release are derived from the same pool of release-ready vesicles (Lu and Trussell, 2000; Otsu et al., 2004). We thus hypothesize that synaptic vesicles normally occupy slots from which they are stimulated for release by Ca 2+. These slots are frozen, at least partially, if synaptotagmin 1 does not become Ca 2+ bound. A possible physical basis for the frozen release slots is the Ca 2+ -independent complex that synaptotagmin 1 forms with SNARE proteins (Bennett et al., 1992; Shin et al., 2003; Rickman and Davletov, 2003). Finally, how can we reconcile our results with studies that report a general increase in asynchronous release in synaptotagmin 1-deficient synapses and conclude that synaptotagmin 1 simply sychronizes release (Yoshihara and Littleton, 2002; Nishiki and Augustine, 2004)? In autaptic neurons, the presynaptic and the postsynaptic cell are one and the same. Thus, even isolated action potentials may result in global cellular Ca 2+ increases that resemble the increase in bulk Ca 2+ observed in interneuronal synapses after stimulus trains, and the type of asynchronous release observed in autapses may effectively correspond to delayed release. Moreover, in Drosophila the major evidence that synaptotagmin 1 synchronizes release was based on the observation that in synaptotagmin 1 mutants, depolarization of nerve terminals in a solution with a low extracellular Ca 2+ concentration (too low to trigger synchronous release) leads to increased asynchronous release in synaptotagmin 1-deficient neurons (Yoshihara and Littleton, 2002). However, this observation exactly duplicates the conditions of delayed release under which we also show that the synaptotagmin 1 deletion indeed potentiates asynchronous release and thus is consistent with the model proposed above. Experimental Procedures Neuron Culture Primary cortical neurons were isolated from E18 or P1 pups of wildtype or synaptotagmin 1-deficient mice, dissociated by trypsin digestion, and plated on Poly-D-lysine-coated glass coverslips (Kavalali et al., 1999). The neurons were cultured in vitro for 12 16 days in MEM (Gibco) supplemented with B27 (Gibco), glucose, transferrin, fetal bovine serum, and Ara-C (Sigma). Electrophysiology Synaptic responses were triggered by 1 ms current injection (900 ma) through a local extracellular electrode (FHC, Inc.) and recorded in a whole-cell mode using Multiclamp 700A amplifier (Axon Instruments, Inc.). The frequency, duration, and magnitude of extracellular stimulus were controlled with Model 2100 Isolated Pulse Stimulator (A-M Systems, Inc.). The whole-cell pipette solution contained 135 mm CsCl 2, 10 mm HEPES, 1 mm EGTA, 1 mm Na-GTP, and 1 mm QX-314 (ph, 7.4). The bath solution contained 140 mm NaCl, 5 mm KCl, 2 mm or 10 mm CaCl 2, 0.8 mm MgCl 2, 10 mm HEPES, and 10 mm glucose (ph, 7.4). Excitatory and inhibitory postsynaptic currents were separated pharmacologically by the addition of 100 mm picrotoxin or 50 mm AP-5 and 20 mm CNQX, respectively, to the bath solution. EGTA-AM (Molecular Probes) was used at a concentration of 100 mm. All further details of the experimental procedures are described in the Supplemental Data. Supplemental Data Supplemental Data include Supplemental Experimental Procedures and seven figures with legends and can be found with this article online at http://www.neuron.org/cgi/content/full/48/4/547/dc1. Acknowledgments We would like to thank Ms. I. Kornblum and A. Roth, for technical assistance, and Drs. C. Rosenmund (Baylor College of Medicine) and Ege Kavalali (UT Southwestern Medical Center), for expert advice. Received: March 11, 2005 Revised: August 16, 2005 Accepted: September 2, 2005 Published: November 22, 2005

Neuron 554 References Atluri, P.P., and Regehr, W.G. (1998). Delayed release of neurotransmitter from cerebellar granule cells. J. Neurosci. 18, 8214 8227. Atwood, H.L., and Karunanithi, S. (2002). Diversification of synaptic strength: presynaptic elements. Nat. Rev. Neurosci. 3, 497 516. Barrett, E.F., and Stevens, C.F. (1972). The kinetics of transmitter release at the frog neuromuscular junction. J. Physiol. 227, 691 708. Bennett, M.K., Calakos, N., and Scheller, R.H. (1992). Syntaxin: a synaptic protein implicated in docking of synaptic vesicles at presynaptic active zones. Science 257, 255 259. Cummings, D.D., Wilcox, K.S., and Dichter, M.A. (1996). Calciumdependent paired-pulse facilitation of miniature EPSC frequency accompanies depression of EPSCs at hippocampal synapses in culture. J. Neurosci. 16, 5312 5323. DiAntonio, A., Parfitt, K.D., and Schwarz, T.L. (1993). Synaptic transmission persists in synaptotagmin mutants of Drosophila. Cell 73, 1281 1290. Fernández-Chacón, R., Königstorfer, A., Gerber, S.H., García, J., Matos, M.F., Stevens, C.F., Brose, N., Rizo, J., Rosenmund, C., and Südhof, T.C. (2001). Synaptotagmin I functions as a Ca 2+ - regulator of release probability. Nature 410, 41 49. Geppert, M., Goda, Y., Hammer, R.E., Li, C., Rosahl, T.W., Stevens, C.F., and Südhof, T.C. (1994). Synaptotagmin I: A major Ca 2+ sensor for transmitter release at a central synapse. Cell 79, 717 727. Goda, Y., and Stevens, C.F. (1994). Two components of transmitter release at a central synapse. Proc. Natl. Acad. Sci. USA 91, 12942 12946. Hagler, D.J., Jr., and Goda, Y. (2001). Properties of synchronous and asynchronous release during pulse train depression in cultured hippocampal neurons. J. Neurophysiol. 85, 2324 2334. Jensen, K., Lambert, J.D., and Jensen, M.S. (2000). Tetanus-induced asynchronous GABA release in cultured hippocampal neurons. Brain Res. 880, 198 201. Kavalali, E.T., Klingauf, J., and Tsien, R.W. (1999). Activitydependent regulation of synaptic clustering in a hippocampal culture system. Proc. Natl. Acad. Sci. USA 96, 12893 12900. Kirischuk, S., and Grantyn, R. (2003). Intraterminal Ca 2+ concentration and asynchronous transmitter release at single GABAergic boutons in rat collicular cultures. J. Physiol. 548, 753 764. Littleton, J.T., Stern, M., Schulze, K., Perin, M., and Bellen, H.J. (1993). Mutational analysis of Drosophila synaptotagmin demonstrates its essential role in Ca2+-activated neurotransmitter release. Cell 74, 1125 1134. Llinas, R., Sugimori, M., and Silver, R.B. (1995). The concept of calcium concentration microdomains in synaptic transmission. Neuropharmacology 34, 1443 1451. Lu, T., and Trussell, L.O. (2000). Inhibitory transmission mediated by asynchronous transmitter release. Neuron 26, 683 694. Nishiki, T., and Augustine, G.J. (2004). Synaptotagmin I synchronizes transmitter release in mouse hippocampal neurons. J. Neurosci. 24, 6127 6132. Otsu, Y., Shahrezaei, V., Li, B., Raymond, L.A., Delaney, K.R., and Murphy, T.H. (2004). Competition between phasic and asynchronous release for recovered synaptic vesicles at developing hippocampal autaptic synapses. J. Neurosci. 24, 420 433. Perin, M.S., Johnston, P.A., Özcelik, T., Jahn, R., Francke, U., and Südhof, T.C. (1991). Structural and functional conservation of synaptotagmin (p65) in Drosophila and humans. J. Biol. Chem. 266, 615 622. Rickman, C., and Davletov, B. (2003). Mechanism of calciumindependent synaptotagmin binding to target SNAREs. J. Biol. Chem. 278, 5501 5504. Robinson, I.M., Ranjan, R., and Schwarz, T.L. (2002). Synaptotagmins I and IV promote transmitter release independently of Ca 2+ binding in the C 2 A domain. Nature 418, 336 340. Shin, O.H., Rhee, J.S., Tang, J., Sugita, S., Rosenmund, C., and Südhof, T.C. (2003). Sr 2+ binding to the Ca 2+ binding site of the synaptotagmin 1 C 2 B domain triggers fast exocytosis without stimulating SNARE interactions. Neuron 37, 99 108. Sorensen, J.B., Fernández-Chacón, R., Südhof, T.C., and Neher, E. (2003). Examining synaptotagmin 1 function in dense core vesicle exocytosis under direct control of Ca 2+. J. Gen. Physiol. 122, 265 276. Tokuoka, H., and Goda, Y. (2003). Synaptotagmin in Ca 2+ -dependent exocytosis: dynamic action in a flash. Neuron 22, 521 524. Voets, T., Moser, T., Lund, P.-E., Chow, R.H., Geppert, M., Südhof, T.C., and Neher, E. (2001). Intracellular calcium dependence of large dense-core vesicle exocytosis in the absence of synaptotagmin I. Proc. Natl. Acad. Sci. USA 98, 11680 11685. Yoshihara, M., and Littleton, J.T. (2002). Synaptotagmin I functions as a calcium sensor to synchronize neurotransmitter release. Neuron 36, 897 908. Zucker, R.S., and Regehr, W.G. (2002). Short-term synaptic plasticity. Annu. Rev. Physiol. 64, 355 405.