Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

An Inhibitory Effect of Extracellular Ca2+ on Ca2+-Dependent Exocytosis

  • Wei Xiong ,

    Contributed equally to this work with: Wei Xiong, Tao Liu

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Tao Liu ,

    Contributed equally to this work with: Wei Xiong, Tao Liu

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Yeshi Wang,

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Xiaowei Chen,

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Lei Sun,

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Ning Guo,

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Hui Zheng,

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Lianghong Zheng,

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Martial Ruat,

    Affiliation CNRS, UPR9040, Institut de Neurobiologie Alfred Fessard-IFR 2118, Gif sur Yvette, France

  • Weiping Han,

    Affiliation Laboratory of Metabolic Medicine, Singapore Bioimaging Consortium, Agency for Science, Technology, and Research, Singapore, Singapore

  • Claire Xi Zhang ,

    zzhou@pku.edu.cn (ZZ); clairexz@pku.edu.cn (CXZ)

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

  • Zhuan Zhou

    zzhou@pku.edu.cn (ZZ); clairexz@pku.edu.cn (CXZ)

    Affiliation State Key Laboratory of Biomembrane Engineering and Center for Life Sciences, Institute of Molecular Medicine, Peking University, Beijing, China

Abstract

Aim

Neurotransmitter release is elicited by an elevation of intracellular Ca2+ concentration ([Ca2+]i). The action potential triggers Ca2+ influx through Ca2+ channels which causes local changes of [Ca2+]i for vesicle release. However, any direct role of extracellular Ca2+ (besides Ca2+ influx) on Ca2+-dependent exocytosis remains elusive. Here we set out to investigate this possibility on rat dorsal root ganglion (DRG) neurons and chromaffin cells, widely used models for studying vesicle exocytosis.

Results

Using photolysis of caged Ca2+ and caffeine-induced release of stored Ca2+, we found that extracellular Ca2+ inhibited exocytosis following moderate [Ca2+]i rises (2–3 µM). The IC50 for extracellular Ca2+ inhibition of exocytosis (ECIE) was 1.38 mM and a physiological reduction (∼30%) of extracellular Ca2+ concentration ([Ca2+]o) significantly increased the evoked exocytosis. At the single vesicle level, quantal size and release frequency were also altered by physiological [Ca2+]o. The calcimimetics Mg2+, Cd2+, G418, and neomycin all inhibited exocytosis. The extracellular Ca2+-sensing receptor (CaSR) was not involved because specific drugs and knockdown of CaSR in DRG neurons did not affect ECIE.

Conclusion/Significance

As an extension of the classic Ca2+ hypothesis of synaptic release, physiological levels of extracellular Ca2+ play dual roles in evoked exocytosis by providing a source of Ca2+ influx, and by directly regulating quantal size and release probability in neuronal cells.

Introduction

Neurotransmitter and hormone secretion are precisely controlled in neurons and endocrine cells where Ca2+ plays a pivotal role [1][3]. In response to action potentials (APs), the entry of extracellular Ca2+ through presynaptic Ca2+ channels results in microdomains of spatial-temporal [Ca2+]i rise which triggers synaptic transmission [2]. At the same time, the rapid influx of Ca2+ through presynaptic Ca2+ channels and then postsynaptic Ca2+-permeable channels also leads to a reduction of extracellular Ca2+ concentration ([Ca2+]o) given the limited Ca2+ storage capacity within the synaptic cleft [4], [5].

Early studies using ion-sensitive electrodes showed that a local electrical stimulus decreases [Ca2+]o by around 1/3 in cerebellar cortex [4]. Direct depolarization of postsynaptic membrane induces Ca2+ influx into the membrane and depletion of Ca2+ in the synaptic cleft. Presynaptic Ca2+ current is thus decreased by 30%, which corresponds to the more than 33% [Ca2+]o drop in a glutamatergic synapse [5]. Under pathological and injury conditions, [Ca2+]o decreases even down to 0.1–0.3 mM from 1.2 mM [6]. It is known that extracellular Ca2+ modulates ion channels in neurons; for example, lowering [Ca2+]o shifts the voltage dependence of sodium channels to more negative potentials, via a biophysical surface charge effect [7]. In addition to this biophysical effect, [Ca2+]o may have a biochemical effect on cell function by regulating non-selective cation channels and TRPM7 channel [8][11].

For neurotransmitter release, it is well established that an AP-induced [Ca2+]i rise directly triggers vesicle release [12]. However, it remains elusive whether extracellular Ca2+ also plays a direct role in vesicle exocytosis from outside cells. To investigate this possibility, we triggered vesicle release using photolysis of caged Ca2+ or caffeine-induced release of stored Ca2+ in the absence of membrane depolarization (without Ca2+ influx through voltage-dependent Ca2+ channels) at various levels of [Ca2+]o. We found that extracellular Ca2+ directly regulates [Ca2+]i-dependent vesicle exocytosis in the somata of sensory dorsal root ganglion (DRG) neurons and neuroendocrine chromaffin cells. Interestingly, physiological levels of [Ca2+]o inhibited [Ca2+]i-dependent exocytosis, partially by altering the quantal size and frequency of single vesicle release.

Methods

Cell preparation and patch clamp recordings

The use and care of animals in this study was approved and overseen by the Institutional Animal Care and Use Committee of Peking University. The permit number for this project is IMM-zhouz-11. Freshly isolated DRG neurons from 120–150 g Wistar rats were prepared as described previously [13], [14]. Cells were used 1–8 h after preparation. Only small neurons (15–25 µm, C-fiber) without apparent processes were selected for experiments.

Rat adrenal medulla slices were prepared as described previously [15]. Briefly, the adrenal glands were removed from adult Wistar rats of 250–300 g. The glands were immediately immersed in ice-cold, low-calcium Ringer's saline (in mM: 125 NaCl, 2.5 KCl, 0.1 CaCl2, 5 MgCl2, 1.25 NaH2PO4, 26 NaHCO3, 12 glucose, pH 7.4) gassed with 95%O2/5%CO2. Single glands were glued with cyanoacrylate to the stage of a vibratome chamber and covered with the same Ringer's saline. Slices of 200–300 µm were cut parallel to the larger base of the gland (Vibrotome 1000, St. Louis, MO). The slices were incubated for 30 min at room temperature in normal Ringer's saline (in mM: 125 NaCl, 2.5 KCl, 2.5 CaCl2, 1 MgCl2, 1.25 NaH2PO4, 26 NaHCO3, 12 glucose) bubbled with 95%O2/5%CO2. The slices were used for up to 8 h after cutting. For amperometric recordings, slices were transferred to a recording chamber attached to the stage of an upright microscope and continuously superfused with Ringer's saline at room temperature (∼22°C) [15].

Standard external medium contained (in mM) 150 NaCl, 5 KCl, 2.5 CaCl2, 1 MgCl2, 10 HEPES, 10 glucose, pH 7.4. Ca2+-free medium was the same except that CaCl2 was removed and 1 mM EGTA was added. Ca2+ and Mg2+ concentrations were changed with osmotic compensation. For solutions with Ca2+ and Mg2+ concentrations between 1 and 100 µM, EGTA was added for Ca2+ and EDTA for Mg2+ to obtain accurate concentrations. Cells were initially bathed in normal external solution and low Ca2+, Ca2+-free solutions, and drugs were locally puffed onto cells using a multichannel microperfusion system (MPS–2; INBIO, Wuhan, China) [16], [17]. Standard intracellular pipette solution contained (in mM) 153 CsCl, 1 MgCl2, 10 HEPES, 4 Mg-ATP, pH 7.2. For flash photolysis experiments, the intracellular pipette solution contained (in mM) 110 Cs-glutamate, 5 NP-EGTA, 8 NaCl, 2.5 CaCl2, 2 Mg-ATP, 0.3 GTP, 0.2 fura-6F, and 35 HEPES, pH 7.2. All chemicals were from Sigma (St. Louis, Missouri), except for NP-EGTA (5 mM), fura-2 AM (final concentration 15 µM), fura-2 salt (20 µM), fura-6F (0.2 mM), and fluo-4 AM (15 µM) (Molecular Probes, Eugene, Oregon).

Membrane capacitance (Cm) measurements

Cm was measured using a software lock-in module of Pulse 8.76 together with an EPC10/2 amplifier, as described previously [14]. Briefly, a 1 kHz, 20 mV peak-to-peak sinusoid was applied around a DC holding potential of −70 mV. The resulting current was analyzed using the Lindau-Neher technique to estimate Cm, membrane conductance and series resistance [18]. Flash-induced maximum capacitance changes were measured as ΔCm.

Amperometry

Highly sensitive, low-noise, 5-µm diameter carbon fiber electrodes (ProCFE, Dagan, Minneapolis, MN) were used for electrochemical monitoring of quantal release of catecholamine from chromaffin cells in slices [19]. Long-tip CFEs with 200 µm sensor tips were used to detect local catecholamine release from many cells [15]. Amperometric currents were recorded at +780 mV, a potential sufficient to oxidize catecholamines [20], [21], low-pass filtered at 300 Hz, and sampled by the EPC-9/2 at 5 kHz.

Flash photolysis of caged Ca2+ and [Ca2+]i measurements

For measurements of [Ca2+]i and UV flashes in DRG cells, we used an IX-71 inverted microscope (Olympus, Tokyo, Japan) equipped with a monochromator-based system (TILL Photonics, Planegg, Germany). NP-EGTA is very selective for Ca2+ and intracellular dialysis of NP-EGTA generated a small loading transient, which did not affect membrane capacitance (data not shown) [3], [22], [23]. The Ca2+ indicator fura-6F was excited at 1 Hz with 5-ms light pulses at 340 and 380 nm for the measurement of [Ca2+]i in the presence of NP-EGTA [24]. The excitation intensity was far below the level for evident photolysis of NP-EGTA. The emitted fluorescence was detected by a photomultiplier tube (TILL), sampled by EPC10/2, and acquired with a Fura extension of the Pulse software (HEKA). UV light from a xenon arc flash lamp (Rapp, Hamburg, Germany) was combined with the excitation light provided by a monochromator before entering an epifluorescence port of the IX-71 microscope. For all flash experiments we used an Olympus Uapo/340 40×, 1.35 NA oil-immersion objective. The calibration method used for caged Ca2+ experiments was similar to that reported [24]. The first flash was triggered 2–3 min after the whole-cell configuration was established. This allowed recovery of [Ca2+]i to ∼300 nM after the loading transient.

For measurements of [Ca2+]i in cells in slices, we used the Uniblitz shutter-based fluorescence system (Vincent, Rochester, NY), a DVC1412 CCD camera (DVC, Austin, TX), and a BX-51 upright microscope (Olympus). The images were acquired with TILLvisION software (TILL Photonics, Planegg, Germany). The adrenal slice was loaded with fluo-4 AM or fura-2 AM by local puffer for ∼120 s via a patch pipette tip of ∼5 µm diameter. Individual cells were imaged by DIC optics using a 40× water-immersion lens (NA = 0.8). For fluo-4, Ca2+ transients were recorded by measuring the emission at 525 nm, resulting from excitation at 488 nm [25]. For fura-2, Ca2+ transients were recorded by measuring the ratio of emission at 510 nm, resulting from excitation at 340 and 380 nm [16].

CaSR knockdown by shRNAs

shRNAs for rat CaSR were designed using Invitrogen BLOCK-iT™ RNAi Designer (https://rnaidesigner.invitrogen.com/rnaiexpress/). Knockdown efficiency was first tested in HEK 293 cells using over-expressed rat CaSR [26]. Five shRNAs were designed and the most efficient two were chosen for knockdown experiments in DRG neurons. The two shRNAs were: sh 1, 5′ -GATCCGCAGGCTCCTCAGCAATAACGAATTATTGCTGAGGAGCCTGCTTTTTTGAATTCA- 3′, and sh 2, 5′ -GATCCGCGCATGCCCTACAAGATATACGAATATATCTTGTAGGGCATGCGCTTTTTTGAATTCA- 3′. Electrophysiology was done 4–5 days after shRNA transfection.

For Western blot, HEK293 cells were lysed in 20 mM HEPES, 1 mM EDTA, 0.1 g/L PMSF, 100 mM NaCl, 1% NP40, pH 7.4, with protease inhibitor cocktail (Calbiochem, San Diego, USA) [27], [28]. Proteins were solubilized in SDS-PAGE buffer. After running SDS-PAGE, they were analyzed by immunoblotting with antibodies to CaSR and β-actin. CaSR antibody was from Affinity BioReagents (Colorado, USA). β-actin antibody was from Sigma (St. Louis, Missouri).

Immunoprecipitations

DRGs were dissected and lysed in 20 mM HEPES, 1 mM EDTA, 0.1 g/L PMSF, 100 mM NaCl, 1% NP40, pH 7.4, with protease inhibitor cocktail (Calbiochem, San Diego, USA) [27], [28]. The complexin antibody was incubated with DRG extract overnight before Protein A beads (GE Healthcare, Uppsala, Sweden) were added. Bound proteins were solubilized in SDS-PAGE buffer. After running SDS-PAGE, they were analyzed by immunoblotting with monoclonal antibodies to syntaxin 1, SNAP 25, synaptotagmin 1, GAPDH and CaSR, and a polyclonal antibody to complexin 2.

Syntaxin 1, SNAP 25, and complexin 2 antibodies were from Synaptic Systems (Göttingen, Germany). Synaptotagmin 1 antibody was from Stressgen Bioreagents (British Columbia, Canada).

Statistics

All experiments were carried out at room temperature (22–25°C), and data are presented as mean ± SEM. The peak values for evoked ΔCm signals were used for analysis, and cell numbers are shown as “n” in brackets. Data were analyzed using Igor software (Wavemetrics, Lake Oswego, Oregon). Significance of difference was tested using either Student's t-test or ANOVA followed by post hoc tests (*p<0.05, **p<0.01, ***p<0.001, n.s. = not significant (p>0.05)).

Results

Extracellular Ca2+ inhibited somatic exocytosis in DRG neurons

DRG neurons are primary sensory cells that release pain-related transmitters including CGRP, substance P and ATP from their terminals and somata in response to APs [13], [16], [29]. There are two kinds of depolarization-induced exocytosis in the somata of DRG neurons [16], [30]; in the present study, we focused on the Ca2+-dependent exocytosis by holding the soma at resting potential [13], [16], [30].

By raising [Ca2+]i through photolysis of caged Ca2+ [24] or caffeine-induced store release [31], we determined whether extracellular Ca2+ directly regulates the [Ca2+]i-dependent exocytosis. We did not use APs, because AP-induced Ca2+ influx causes microdomain changes in [Ca2+]i which cannot be controlled under various levels of [Ca2+]o [5], [32]. Isolated DRG neurons underwent whole-cell dialysis with 5 mM of the caged-Ca2+ compound nitrophenyl-EGTA. We elicited a [Ca2+]i rise using photolysis to a level near the threshold of evoked exocytosis. A train of low-energy UV flashes was applied to produce a [Ca2+]i plateau of ∼2 µM, resulting in a large Cm increase of ∼2 pF in a Ca2+-free bath (Figure 1A). This Cm increase reflects a substantial amount of exocytosis (∼4000 vesicles, assuming 0.5 fF per 140 nm vesicle [16]). However, 2 min after adding 2.5 mM Ca2+ to the bath, similar flash-induced [Ca2+]i transients nearly failed to induce Cm increases in the same neuron (Figure 1A). The evoked ΔCm was fully recovered after removing external Ca2+. In this particular cell, physiological levels of extracellular Ca2+ (2.5 mM) fully inhibited the [Ca2+]i-dependent ΔCm. In a second cell, the UV-flash evoked a similar level of [Ca2+]i rise and ΔCm in 2.5 mM [Ca2+]o was inhibited again, albeit by 40% compared to the Ca2+-free bath (Figure 1B). These two cells represent the maximal and minimal inhibition we observed. To compare the rates of photolysis-induced exocytosis in normal external Ca2+ solution and Ca2+-free solution, we averaged ΔCm responses within 9 s after photolysis. Photolysis induced a much larger ΔCm in Ca2+-free solution (Figure 1C, upper panel). We then used normalized peak ΔCm to quantify extracellular Ca2+ inhibition of exocytosis (ECIE). Compared to 0 mM [Ca2+]o, the evoked [Ca2+]i rise was similar (106±8%), but ΔCm was reduced by 82±8% in 2.5 mM [Ca2+]o (Figure 1C, lower panel). Similar ECIE was also found when a [Ca2+]i rise was elicited by caffeine, which releases Ca2+ from the intracellular ryanodine-sensitive Ca2+ store [31] (Figure 1D). The elicited [Ca2+]i was nearly identical, while ΔCm was reduced by 58±5% in 2.5 mM [Ca2+]o (Figure 1E). Thus, extracellular Ca2+ inhibited exocytosis in the somata of rat DRG neurons.

thumbnail
Figure 1. Extracellular Ca2+ inhibited [Ca2+]i-dependent exocytosis in DRG neurons.

(A) Membrane capacitance responses to Ca2+ spikes induced by photolysis in the presence (2.5 mM) or absence (0 mM) of external Ca2+. The internal solution in the whole-cell recording pipette contained the photolabile Ca2+ chelator nitrophenyl-EGTA. The arrowheads mark the application of four UV flashes at 0.5 Hz. In one DRG neuron, from the top, individual traces show changes in membrane capacitance (ΔCm), series resistance (Gs), membrane conductance (Gm), and [Ca2+]i. There were 2-min breaks between recordings. The initial values of Cm, Gs, Gm, and Im are noted. [Ca2+]i was monitored by Fura-6F measurements. (B) In a second DRG neuron, for clarity, only Cm and [Ca2+]i are shown. (C) Upper panel: averaged Cm responses in 0 and 2.5 mM [Ca2+]o within the first 9 s after photolysis (n = 10). Averaged initial Cm was 21.4±1.2 pF. Lower panel: statistics of normalized peak ΔCm and [Ca2+]i following photolysis. Compared to 0 mM [Ca2+]o, the evoked [Ca2+]i rise was similar (106±8%), but exocytosis was reduced by 82±8% in 2.5 mM [Ca2+]o (n = 10, p<0.001). (D) Membrane capacitance responses to Ca2+ rise induced by local caffeine application (caff, 20 mM) in the presence (2.5 mM) or absence (0 mM) of external Ca2+ in a DRG neuron. There were 90-s breaks between recordings. (E) Normalized peak ΔCm and [Ca2+]i following caffeine stimulation. Compared to 0 mM [Ca2+]o, the [Ca2+]i rise was similar, while exocytosis was reduced by 58±5% in 2.5 mM [Ca2+]o (n = 14, p<0.001).

https://doi.org/10.1371/journal.pone.0024573.g001

Exocytosis at both 0 and 2.5 mM [Ca2+]o was facilitated by cAMP elevation with forskolin, presumably via activation of protein kinase A (PKA), which is a well-known feature of Ca2+-dependent exocytosis for synaptic transmission in hippocampal neurons and somatic release in chromaffin cells (Figure S1) [33]. Thus, the ECIE-targeted vesicles use the classic type of Ca2+-dependent exocytosis.

Because membrane capacitance measures the sum of membrane-expanding exocytosis and membrane-retrieving endocytosis, and a 0 Ca2+ external solution may arrest the internalization of vesicles [34], the lack of a Cm response to UV flashes at 2.5 mM [Ca2+]o could be due to a similar-sized endocytosis. We examined this possibility by inhibiting dynamin, a GTPase essential for most endocytotic pathways [35]. Endocytosis was greatly accelerated when the external solution was shifted to 2.5 mM Ca2+ from 0 Ca2+ (Figure 2A, right panel), which corresponds with a previous report [34]. Intracellular dialysis of GTPγS (200 µM) inhibited endocytosis without affecting the ECIE in DRG neurons (Figure 2B–D). Note that GTPγS also reduced partial exocytosis in DRG neurons (Figure 2), implying that GTP affected exocytosis in these neurons [36], [37]. Since endocytosis was increased in 2.5 mM Ca2+ (Figure 2A), ECIE measured by Cm could be partially contaminated by exocytosis-coupled endocytosis.

thumbnail
Figure 2. GTPγS had no effect on ECIE in DRG neurons.

(A) Responses of ΔCm and [Ca2+]i in a control neuron without GTPγS. The peak of flash-induced exocytosis is Cm1 while Cm2 represents a robust endocytosis in 40 s. The ratio Cm2/Cm1 (Rendo/exo) reflects the degree of endocytosis compared to the preceding exocytosis [14]. The initial value of Cm is noted. [Ca2+]i was monitored by Fura-6F measurements. Note the shift of extracellular solution from 0 Ca2+ to 2.5 mM Ca2+ accelerated endocytosis. (B) A different neuron dialyzed with 200 µM GTPγS, where exocytosis was induced only in the absence of external Ca2+. However, following the evoked exocytosis in 0 Ca2+, there was little decline in Cm (endocytosis) when the external bath was changed to 2.5 mM Ca2+. (C) Average results of ECIE in the presence of GTPγS. Compared with 0 Ca2+, ΔCm in 2.5 mM Ca2+ was reduced to 18±14% (n = 5, p<0.001), while [Ca2+]i was not reduced (121±13%). (D) Average ratios of the evoked endocytosis and exocytosis (Rendo/exo). Rendo/exo was 45±8% with GTPγS but 136±8% without GTPγS (n = 6, p<0.001).

https://doi.org/10.1371/journal.pone.0024573.g002

To confirm the ECIE with a non-capacitance assay and investigate its role in transmitter release, we used a neuroendocrine cell which is a widely used model for exocytosis [31], [38], [39], the rat adrenal chromaffin cell (RACC). Combining membrane capacitance and amperometric recording of catecholamine release, we found that ECIE occurred in chromaffin cells too. Compared to photolysis-induced exocytosis in Ca2+-free solution, the photolysis-induced ΔCm was inhibited to 69±5%, the number of amperometric spikes was reduced to 68±6%, and the integrated amperometric signal was inhibited to 59±9% (Figure S2). The photolysis-induced [Ca2+]i was not changed (107±4%). Taken together, extracellular Ca2+ inhibited Ca2+-regulated vesicle exocytosis.

Attenuation of ECIE by high [Ca2+]i rise

The above experiments showed ECIE in moderate [Ca2+]i (2–3 µM). These data are not consistent with other studies using photolysis of caged Ca2+ to much higher [Ca2+]i levels (6–25 µM) [24], [40], [41]. Thus, we investigated ECIE at higher [Ca2+]i by using high-energy UV flashes and 10 mM nitrophenyl-EGTA. Extracellular Ca2+ still reversibly inhibited exocytosis when the UV flash-induced [Ca2+]i rise was increased to ∼4 µM (Figure 3A). Statistically, the evoked exocytosis at [Ca2+]o of 2.5 mM was inhibited by 40±5% of that at a [Ca2+]o of 0 (Figure 3B), much less than the 82±8% inhibition at a 2–3 µM [Ca2+]i rise, indicating that ECIE was attenuated by higher [Ca2+]i rises (p<0.001, Figure 3C, Figure S3). Therefore, the inhibition by [Ca2+]o could be overcome by higher [Ca2+]i.

thumbnail
Figure 3. High [Ca2+]i attenuated ECIE.

(A) ΔCm and [Ca2+]i signals in response to Ca2+ release by trains of UV flashes at high energy. The initial value of Cm is noted. [Ca2+]i was monitored by Fura-6F measurements. (B) Statistics of normalized ΔCm and [Ca2+]i (n = 14, p<0.001). UV flash-induced [Ca2+]i rises were similar, 6.3±0.4 µM ([Ca2+]o = 0) and 6.4±0.4 µM ([Ca2+]o = 2.5 mM). (C) Comparison of ECIE (or normalized exocytosis ΔCm) in high and low [Ca2+]i. Compared to 0 mM [Ca2+]o, evoked exocytosis was reduced by 82±8% ([Ca2+]i = 2 µM, n = 10) and 40±5% ([Ca2+]i = 6 µM, n = 14) in 2.5 mM [Ca2+]o.

https://doi.org/10.1371/journal.pone.0024573.g003

Changes of [Ca2+]o within the physiological range modulated exocytosis

[Ca2+]o varies during normal physiological neural activity [4], [5], [42]. Direct depolarization of the postsynaptic membrane causes a more than 33% drop of [Ca2+]o in a glutamatergic synapse [5]. In single DRG neurons, we reduced [Ca2+]o 34% (from 2.5 mM to 1.65 mM) and tested its effect on evoked exocytosis. The caffeine-induced exocytosis was 35% greater in 1.65 mM than in 2.5 mM [Ca2+]o, while the evoked [Ca2+]i rise was similar (Figure 4). Thus, these experiments strongly suggested that ECIE occurred under physiological conditions in DRG neurons.

thumbnail
Figure 4. Physiological levels of extracellular Ca2+ decrease modulated exocytosis in DRG neurons.

(A) Combined ΔCm and [Ca2+]i recordings in a DRG neuron. A reduction of [Ca2+]o from 2.5 mM to 1.65 mM (34%) increased the caffeine-induced (20 mM) ΔCm signal. Upper, ΔCm signals following caffeine stimulation. Lower, corresponding [Ca2+]i rise monitored by Fura-2 measurements. There were 90-s breaks between recordings. The initial value of Cm is noted. AU, arbitrary units. (B) Statistics of normalized exocytosis and [Ca2+]i rise signals. A 34% reduction of [Ca2+]o increased caffeine-induced exocytosis signals by 35±8% (n = 9, p<0.01). The caffeine-induced [Ca2+]i rise was similar (99±2%, n = 8) when [Ca2+]o was changed between 1.65 mM and 2.5 mM.

https://doi.org/10.1371/journal.pone.0024573.g004

Calcimimetics inhibited [Ca2+]i-dependent exocytosis in DRG neurons

In search of the molecular mechanism underlying ECIE, we investigated the extracellular calcium-sensing receptor (CaSR), which inhibits secretion in parathyroid cells [43]. The CaSR is expressed in nerve terminals [26] and in DRG neurons [44], [45]. First, we used CaSR agonists (or “calcimimetics”), the divalent ions, Ca2+, Mg2+, and Cd2+, G418, and neomycin [43], [46]. Similar to Ca2+, G418 or other calcimimetics also inhibited caffeine-induced exocytosis in DRG neurons, while the caffeine-induced [Ca2+]i rise was unaffected (Figure 5A). In these experiments, G418 was the most potent inhibitor with an EC50 of 28 µM, followed by Cd2+ (47 µM), Mg2+ (1.26 mM), and Ca2+ (1.38 mM) (Figure 5B). Second, since these calcimimetics may not be specific to the CaSR [9], [47], we assessed the effects of the more specific CaSR antagonist calhex 231 (calhex) and the agonist calindol [48], [49]. In HEK 293 cells expressing rat CaSR, calhex inhibited the [Ca2+]i rise induced by high [Ca2+]o perfusion while calindol increased [Ca2+]i (Figure S4). However, calhex failed to affect caffeine-induced exocytosis in 2.5 mM [Ca2+]o and calindol did not affect exocytosis in 0 mM [Ca2+]o (Figure 5C, D), or in 0.5 mM [Ca2+]o (data not shown). Neither calhex (1 µM) nor calindol (1 µM) affected the caffeine-induced [Ca2+]i rise in DRG neurons (Figure S5). Third, in cortical synaptosomes, a non-selective cation channel (NSCC) coupled with the CaSR is modulated by [Ca2+]o [9]. NSCC may modulate synaptic transmissions in cortical neurons [50]. We determined whether NSCC existed in DRG neurons and found that it was not detectable (Figure S6). Fourth, we determined whether the CaSR was physically linked to the SNARE complex, which is responsible for all vesicle exocytosis [51]. Co-immunoprecipitation with complexin failed to detect the CaSR in the presence or absence of 3.5 mM Ca2+, while syntaxin1 and SNAP25 were found (Figure S7) as reported in brain [52].

thumbnail
Figure 5. Characterization of ECIE.

(A) Combined ΔCm and [Ca2+]i measurements in a DRG neuron. A representative neuron showing that G418 inhibited the exocytosis induced by caffeine (20 mM). Upper traces: ΔCm responses to application solutions containing caffeine and G418 (0.8 mM). The initial value of Cm is noted. Lower traces: corresponding [Ca2+]i monitored by Fura-2 measurements. There were 90-s breaks between recordings. (B) Dose-dependence of inhibition of exocytosis by [Ca2+]o (EC50 = 1.38 mM, n = 6), [Mg2+]o (EC50 = 1.26 mM, n = 6), [Cd2+]o (EC50 = 47 µM, n = 6), and [G418]o (EC50 = 28 µM, n = 6). (C) Cm responses to application solutions containing caffeine and calhex (1 µM) or calindol (1 µM). (D) Statistics of the caffeine-induced ΔCm signals. Compared to exocytosis in 0 mM Ca2+ and 2.5 mM Ca2+, caffeine-induced ΔCm was 98±1% in calindol (n = 7) and 107±3% in calhex (n = 6).

https://doi.org/10.1371/journal.pone.0024573.g005

Finally, we undertook a RNAi knockdown approach to test whether the CaSR was responsible for ECIE. Two short hairpin RNAs (shRNAs) for rat CaSR were designed and tested in HEK 293 cells. Both shRNAs efficiently knocked down over-expressed rat CaSR (Figure S8A). However, CaSR knockdown in DRG neurons had no effect on ECIE (Figure S8B, C).

Taken together, we concluded that ECIE was not mediated through the CaSR, although the ECIE sensor and CaSR shared agonists of Ca2+ and other divalent ions, as well as their potency order (Figure 5B).

ECIE in quantal transmitter release

Having shown that extracellular Ca2+ inhibited photolysis-induced exocytosis in single DRG neurons and chromaffin cells (Figures 1, 2, and 3, Figure S2), we next tested whether extracellular Ca2+ also inhibited exocytosis in the rat adrenal slice. Following application of caffeine, individual quantal exocytosis of catecholamines was detected by a micro carbon fiber electrode (CFE) placed in an adrenal slice [15], [53] (Figure 6A). Since extracellular Ca2+ and Mg2+ both inhibited exocytosis (Figure 5B), we used Ca2+- and Mg2+-free solution to remove calcimimetic inhibition of catecholamine release. In Ca2+- and Mg2+-free solution the baseline of recording shifted upwards (Figure 6A), probably due to the catecholamine release from distant cells because it was absent when the CFE holding potential was below the oxidization voltage (0 mV, [20], [38], [39]). The caffeine-induced exocytosis was inhibited in 2.5 mM [Ca2+]o (Figure 6A), while the [Ca2+]i rise was similar (Figure 6B). These experiments demonstrated that ECIE occurred in quantal catecholamine transmitter release in RACCs. A physiological reduction (∼34%) of extracellular Ca2+ also significantly increased caffeine-induced transmitter release in RACCs (Figure S9).

thumbnail
Figure 6. ECIE in quantal transmitter release.

(A) Left, upper traces: amperometric recording of typical secretory signals evoked by 30 s caffeine (20 mM) at 0 and 2.5 mM [Ca2+]o in a rat adrenal slice. Recordings were performed at 120-s intervals. Lower traces show the integrated amperometric current signals (∫Iampdt, or total catecholamine molecules detected). Right, average of normalized ∫Iampdt signals from 15 cells. Caffeine-induced exocytosis was reduced to 22±3% in 2.5 mM vs 0 mM [Ca2+]o (n = 15, p<0.001). (B) Left, caffeine-induced [Ca2+]i rise was similar when changing [Ca2+]o in an adrenal slice. [Ca2+]i rise in response to standard 20 mM caffeine at 0 and 2.5 mM [Ca2+]o was measured using fluo-4 AM. Right, average of caffeine-induced [Ca2+]i rises from 20 cells. The caffeine-induced [Ca2+]i rise was similar at 0 and 2.5 mM [Ca2+]o (101±5%, n = 20). (C) The distribution of quantal size of caffeine-induced amperometric spikes in 0 Ca2++0 Mg2+ (black traces, 423 events) and 2.5 mM Ca2++1 mM Mg2+ (gray traces, 150 events) bath solutions. Histograms show the percentage of amperometric quantal sizes. The curves show Gaussian fits of the distributions. Right inset: average events of single amperometric spikes from the top 10% of fastest events in 0 Ca2++0 Mg2+ (n = 13 cells) vs 2.5 mM Ca2++1 mM Mg2+ (n = 7) [31]. (D) Quantal analysis. Data shown are mean ± s.e.m. Left, parameters for analysis. Right, statistical table. (Q, quantal size; PA, peak amplitude; HHD, half-height duration; RT, rise-time; f, normalized frequency).

https://doi.org/10.1371/journal.pone.0024573.g006

To investigate the extracellular Ca2+ effect on single vesicle release kinetics, we undertook a quantitative analysis of amperometric spikes representing individual vesicle release. Several features of the spikes are used to reflect the kinetics of vesicle fusion (Figure 6D, left panel) [31], [38], [54]. Compared to caffeine-induced amperometric spikes in bath with 0 Ca2++0 Mg2+, bath containing 2.5 mM Ca2++1 mM Mg2+ inhibited the quantal size (Q) by ∼27% (Figure 6D, table) and shifted the Q distribution towards smaller sizes (Figure 6C). In addition, the peak amplitude was inhibited by ∼47%, while half-height duration and rise time were increased by ∼30% and ∼50%, and the normalized frequency of the amperometric spikes was reduced by ∼60% (Figure 6D, table). These changes in amperometric signals are different from (but not in conflict with) a previous report where no change in quantal size was found under non-physiological high [Ca2+]o (5–90 mM) (compared to the physiological 2.5 mM) [55].

Discussion

In this work we used photolysis of a caged Ca2+ compound or caffeine to elevate [Ca2+]i without Ca2+ influx. This permitted us to determine whether extracellular Ca2+ directly regulated vesicle release. We found that ECIE occurred in the somata of DRG neurons and adrenal chromaffin cells.

ECIE is supported by the following evidence: (1) for a given rise in [Ca2+]i induced by UV flash photolysis, the [Ca2+]i-induced Cm increase was enhanced by removing external Ca2+ in DRG neurons (Figure 1); (2) intracellular dialysis of GTPγS, which inhibits exocytosis-coupled endocytosis, had no effect on ECIE (Figure 2, see also [35]), indicating that the reduced ΔCm caused by external Ca2+ was not due to an enhanced dynamin-dependent endocytosis in normal external Ca2+ solution [34]; (3) direct measurements of catecholamine release using amperometry detected ECIE in rat single chromaffin cells (Figure S2) and adrenal slices (Figure 6); (4) extracellular Ca2+ and Mg2+ reduced quantal size (by 49%) and release probability (by 60%), and modulated the release kinetics of single vesicle exocytosis (Figure 6C, D); (5) physiological reduction of [Ca2+]o (∼34%) significantly increased exocytosis in DRG neurons (Figure 4) and adrenal slices (Figure S9); (6) the exocytotic inhibition by [Ca2+]o was dose-dependent (Figure 5B); and (7) ECIE was not caused by a Ca2+ gradient near the plasma membrane, which might be produced by possible Ca2+ efflux following a sudden [Ca2+]i rise, because Ca2+ efflux during [Ca2+]i-induced exocytosis would only reduce exocytosis.

ECIE was induced by physiological changes in [Ca2+]o. A 34% change (from 2.5 to 1.65 mM) [4], [5], [56], [57] of [Ca2+]o significantly increased total exocytosis (Figure 4 and Figure S9). Thus, a Ca2+ influx via Ca2+-permeable channels has dual effects on exocytosis: increasing [Ca2+]i and depleting [Ca2+]o [5], [42], [58], both of which boost exocytosis. Ca2+ may affect fusion pore/vesicle exocytosis from both sides of the plasma membrane: from the well-known intracellular site via Ca2+-sensing proteins, and from the as yet unknown extracellular site of the ECIE. To interpret less ECIE at higher [Ca2+]i, we hypothesize (1) the opening of a fusion pore is contributed to by both an intracellular force (Fin by synaptotagmins [51]) and another extracellular force (Fout of ECIE); (2) the total threshold force to open the fusion pore is Ftotal, such that Fin+Fout≥Ftotal is the condition to open fusion pore/exocytosis. This hypothesis explains that increasing Fin by higher [Ca2+]i reduces the effect of [Ca2+]o on ECIE.

ECIE does not at all contradict the well-established Ca2+ hypothesis of vesicle exocytosis [1], [59]. Instead, our data confirmed that the exocytosis was strictly dependent on [Ca2+]i. As an additional effect of [Ca2+]o, ECIE had never been examined previously under proper conditions. Most previous work used membrane depolarization/APs to trigger vesicle release [1], [3], [32], [59]. Synaptic vesicles are docked for exocytosis in active zones where Ca2+ channels are also clustered. AP-triggered vesicle release is induced by Ca2+ microdomains near the Ca2+ channel (at 20 nm, ∼100 µM; at 200 nm, ∼5–10 µM) [32], [60]. ECIE was best seen when a subthreshold [Ca2+]i (2∼3 µM) was triggered (Figure 1). A higher [Ca2+]i (∼6 µM) greatly attenuated this inhibition (Figure 3). Under physiological conditions, AP-induced secretion is achieved by mixed effects of both [Ca2+]i and [Ca2+]o. This explains why extracellular inhibition of exocytosis was overlooked by most previous work.

Regarding mechanisms of ECIE, (1) we investigated whether ECIE is mediated via the CaSR, which is a known [Ca2+]o sensor inhibiting cell secretion in parathyroid cells [43], [61]. Although Ca2+ and other extracellular calcimimetics (Mg2+, Cd2+ and G418) inhibited exocytosis, we failed to link the CaSR to ECIE (Figure 5C, D; Figures S7, S8). An unknown Ca2+-binding protein at the extracellular side of the plasma membrane may be responsible for ECIE. (2) Generally Ca2+-dependent secretion is SNARE-dependent, although did not determine whether ECIE is SNARE dependent, because neurotoxins against somatic secretion in DRG neurons are not available. (3) Conformational changes of the Ca2+ channel induced by membrane depolarization and Ca2+ binding are reported to facilitate vesicle exocytosis in PC12 cells and chromaffin cells [62], [63]. This depolarization effect shares with ECIE the facilitation of exocytosis without Ca2+ influx. However, this effect is different from ECIE since there is no membrane depolarization to open Ca2+ channel for the Ca2+ entry. (4) It is known that extracellular Ca2+ could shifts the voltage-dependent gating of Na+ and Ca2+ channels through a surface charge effect [7], [64]. Future work may determine whether EICE is altered by the surface charge effect.

In summary, our study demonstrated that extracellular Ca2+ directly regulated exocytosis. It is known that under pathological conditions, such as brain ischemia, extracellular Ca2+ and Mg2+ decrease greatly [6], [65]. This may directly enhance synaptic transmission according to our finding. ECIE was most effective at moderate levels (≤6 µM), but less effective in high [Ca2+]i. Thus, ECIE may affect neurotransmitter release more in neural somata than in synapses, where synchronized exocytosis is triggered by Ca2+ microdomains (10∼100 µM) following an AP [32]. ECIE could be a new extension of the classic Ca2+ hypothesis of exocytotic transmitter release [1], [60].

Supporting Information

Figure S1.

Facilitation of photolysis-induced exocytosis by cAMP elevation with forskolin in DRG neurons. (A) ΔCm and [Ca2+]i signals in response to Ca2+ release by trains of UV flashes (4 flashes at 0.5 Hz). The neuron was perfused with standard bath solutions containing 0 mM Ca2+, 0 mM Ca2++100 µM forskolin, 2.5 mM Ca2+, and 2.5 mM Ca2++100 µM forskolin, respectively, at 120 s intervals. UV flashes produced similar [Ca2+]i changes, while the corresponding exocytosis was facilitated by forskolin both in the presence and absence of external Ca2+. The initial value of Cm is noted. [Ca2+]i was monitored by Fura-6F measurements. (B) Comparison of averaged ΔCm traces with forskolin in 0 Ca2+ (left, n = 4) and 2.5 mM Ca2+ (right, n = 4). (C–D) Average results from 4 cells showing that forskolin enhanced exocytosis in both 0 and 2.5 mM Ca2+. Exocytosis in 0 and 2.5 mM Ca2+ in the absence of forskolin was 70±8% and 22±9% of that with forskolin treatment. The corresponding [Ca2+]i values were similar in the presence and absence of forskolin. Compared to the [Ca2+]i rise with forskolin application, the [Ca2+]i rise was 88±6% in 0 [Ca2+]o and 103±5% in 2.5 mM [Ca2+]o solutions.

https://doi.org/10.1371/journal.pone.0024573.s001

(DOC)

Figure S2.

ECIE measured by combined membrane capacitance and amperometry in cultured rat adrenal chromaffin cells. (A) Representative ΔCm and amperometric current (Iamp) responses to [Ca2+]i rises induced by UV flashes in the presence (2.5 mM) or absence (0 mM) of external Ca2+. The arrowheads indicate the time points of UV flashes. A [Ca2+]i increases and secretion signals (ΔCm and Iamp or ∫Iampdt, the oxidization charge proportional to the number of oxidized catecholamines) were first induced by a train of UV flashes in 2.5 mM extracellular Ca2+ solution (2.5 Ca2+) (left). Subsequently, another similar [Ca2+]i increase induced by the second UV train produced much larger secretion signals in Ca2+-free solution (0 Ca2+) (middle). Finally, a similar [Ca2+]i rise induced by the third UV train triggered secretion signals similar to that by the first UV train when extracellular solution was changed back to 2.5 mM (right). (B) Statistics. Following UV flashes, the secretion signals were significantly smaller in 0 vs 2.5 mM Ca2+ (ΔCm, spike numbers and ∫Iampdt were 69±5, 68±6 and 59±9% of controls), while the [Ca2+]i values were similar (107±4% of control).

https://doi.org/10.1371/journal.pone.0024573.s002

(DOC)

Figure S3.

Extracellular Ca2+ did not inhibit high [Ca2+]i-induced exocytosis. Increase of [Ca2+]i to ∼8 µM almost abolished the extracellular Ca2+ inhibition of photolysis-induced exocytosis in DRG neurons.

https://doi.org/10.1371/journal.pone.0024573.s003

(DOC)

Figure S4.

Positive control of specific reagents (calhex and calindol) against CaSR. (A) Typical trace of [Ca2+]i measurement in CaSR-transfected HEK 293 cells. Cells initially bathed in 10 µM [Ca2+]o solution. [Ca2+]o at 1 mM induced a [Ca2+]i rise, which was smaller in the presence of calhex (1 µM) (n = 4). [Ca2+]i was monitored using Fura-2 AM. (B) Cells bathed in 0.25 mM [Ca2+]o solution. Calindol (1 µM) induced a [Ca2+]i rise (n = 5).

https://doi.org/10.1371/journal.pone.0024573.s004

(DOC)

Figure S5.

Calhex and calindol had no effect on [Ca2+]i rise induced by caffeine. (A) [Ca2+]i rise induced by caffeine was monitored by Fura-2 measurements. Left traces show the [Ca2+]i rise induced by 20 mM caffeine in the presence of 2.5 mM [Ca2+]o and 1 µM calhex. Right traces show the [Ca2+]i rise induced by caffeine in the presence of 0 mM [Ca2+]o and 1 µM calindol. [Ca2+]i was monitored by Fura-2 measurements. There was a 90-s interval between continuous recordings. (B) Statistics of the caffeine-induced [Ca2+]i rise. Compared to signals in 0 mM [Ca2+]o and 2.5 mM [Ca2+]o, the [Ca2+]i rise was similar in calindol (97±4%, n = 5) or calhex (100±2%, n = 5).

https://doi.org/10.1371/journal.pone.0024573.s005

(DOC)

Figure S6.

NSCCs did not exist on DRG neurons. DRG neurons were on-cell patched. The upper four traces show consecutive current recordings at 60-s intervals in different perfusion solutions (#1, 2.5 mM Ca2+ and 1 mM Mg2+; #2, 0.1 mM Ca2+ and 0 mM Mg2+; #3, 0 mM Ca2+ and 0 mM Mg2+; #4, 2.5 mM Ca2+ and 1 mM Mg2+). The lowest trace shows the stimulation pulse. DRG neurons were stimulated with a step depolarization from −40 mV to +150 mV, followed by a hyperpolarization to −100 mV. Voltage-induced membrane currents in different external solutions were the same and this excludes the existence of NSCCs in the somata of DRG neurons [9], [50].

https://doi.org/10.1371/journal.pone.0024573.s006

(DOC)

Figure S7.

CaSR is not linked to SNARE complex. (A) Immunoblot of exocytosis-related proteins in DRG extracts with antibodies against syntaxin 1, synaptotagmin 1, SNAP 25, and complexin 1 and 2. GAPDH was used as a control. For the complexin doublet, the upper band is complexin 2 and the lower band is complexin 1. (B) Immunoprecipitation using a polyclonal antibody against aa 45–81 in complexin 2 with or without 3.5 mM Ca2+. Bound proteins were detected with monoclonal antibodies to syntaxin 1, SNAP 25, and CaSR. The upper strip shows that complexin antibody immunoprecipitated syntaxin 1 and SNAP 25. The lower strip shows no CaSR in the complexin immunocomplex.

https://doi.org/10.1371/journal.pone.0024573.s007

(DOC)

Figure S8.

CaSR knockdown had no effect on ECIE. (A) Western blot of CaSR and β-actin in shRNA 1 (sh 1), shRNA 2 (sh 2) and control shRNA-treated HEK 293 cells. Note that the lower band (CaSR) is missing in sh 1 and sh 2-treated cells. The upper band labeled with an asterisk (*) is a non-specific band recognized by the antibody. (B) Electrophysiology of ECIE in control shRNA, sh 1 and sh 2-treated DRG neurons. Experiments were done 4–5 days after transfection. (C) Statistics of normalized exocytosis and [Ca2+]i rise. The caffeine (20 mM)-triggered exocytosis in 2.5 mM Ca2+ was reduced to 41±8% (n = 6), 40±5% (n = 6) and 50±6% (n = 6) in control shRNA, sh 1 and sh 2, respectively. Extracellular Ca2+ inhibition was unaffected in sh 1 and sh 2-treated DRG neurons compared to that of control. [Ca2+]i was monitored by Fura-2 measurements. Compared to the [Ca2+]i rise in 0 mM Ca2+, the caffeine-induced Ca2+ rise was greater in 2.5 mM Ca2+ in the control cells (135±11%, p = 0.03, n = 5). In sh 1 and sh 2-transfected neurons. the caffeine-induced [Ca2+]i rise was similar to that in 0 mM Ca2+ and 2.5 mM Ca2+ (sh 1-tranfected neurons, 112±18%, n = 5; sh 2-transfected neurons, 115±11%, n = 6).

https://doi.org/10.1371/journal.pone.0024573.s008

(DOC)

Figure S9.

Physiological levels of extracellular Ca2+ decrease modulated exocytosis in chromaffin cells. (A) Left, amperometric recording from an adrenal slice. A 34% reduction of [Ca2+]o increased the 20 mM caffeine-induced amperometric signal. Upper traces show amperometric recordings, lower traces show the corresponding ∫Iampdt signals. Right, the caffeine-induced [Ca2+]i rise was similar when changing [Ca2+]o in another rat adrenal slice. [Ca2+]i was measured using Fura-2 AM. (B) Statistics of normalized exocytosis and [Ca2+]i rise signals. A 34% [Ca2+]o reduction increased caffeine-induced exocytosis signals by 68±12% in chromaffin cells (n = 7, p<0.001). The caffeine (20 mM)-induced [Ca2+]i rise was similar at [Ca2+]o of 1.65 mM and 2.5 mM in chromaffin cells (107±7%, n = 15).

https://doi.org/10.1371/journal.pone.0024573.s009

(DOC)

Acknowledgments

We thank Dr. Emmanuel Awumey for advice on DRG extracts, Drs. Tao Xu, Li-Jun Kang and Zi-Xuan He for help with the UV-flash technique, Drs. Juan-Juan Ji and Shu-Yun Bai for Ca2+ imaging in slices, Dr. R. H. Dodd for providing the calhex 231 and calindol, and Drs. Chen Zhang and I.C. Bruce for helpful comments.

Author Contributions

Conceived and designed the experiments: ZZ WPH CXZ. Performed the experiments: WX TL YSW XWC JL XYZ SG. Analyzed the data: WX TL. Contributed reagents/materials/analysis tools: LS NG HZ LHZ MR. Wrote the paper: ZZ WX TL. Calhex and calindo work: TL MR.

References

  1. 1. Katz B (1969) The Release of Neural Transmitter Substances. Liverpool: Liverpool Univerity Press. pp. 1–60.
  2. 2. Augustine GJ, Charlton MP, Smith SJ (1987) Calcium action in synaptic transmitter release. Annu Rev Neurosci 10: 633–693.
  3. 3. Neher E, Zucker RS (1993) Multiple calcium-dependent processes related to secretion in bovine chromaffin cells. Neuron 10: 21–30.
  4. 4. Nicholson C, ten Bruggencate G, Stockle H, Steinberg R (1978) Calcium and potassium changes in extracellular microenvironment of cat cerebellar cortex. J Neurophysiol 41: 1026–1039.
  5. 5. Borst JG, Sakmann B (1999) Depletion of calcium in the synaptic cleft of a calyx-type synapse in the rat brainstem. J Physiol 521 Pt 1: 123–133.
  6. 6. Zhang ET, Hansen AJ, Wieloch T, Lauritzen M (1990) Influence of MK-801 on brain extracellular calcium and potassium activities in severe hypoglycemia. J Cereb Blood Flow Metab 10: 136–139.
  7. 7. Hille B (2001) Ion Channels of Excitable Membranes. Sunderland, Massachusetts, USA: SINAUER ASSOCIATES, INC. pp. 646–651.
  8. 8. Hablitz JJ, Heinemann U, Lux HD (1986) Step reductions in extracellular Ca2+ activate a transient inward current in chick dorsal root ganglion cells. Biophys J 50: 753–757.
  9. 9. Smith SM, Bergsman JB, Harata NC, Scheller RH, Tsien RW (2004) Recordings from single neocortical nerve terminals reveal a nonselective cation channel activated by decreases in extracellular calcium. Neuron 41: 243–256.
  10. 10. Wei WL, Sun HS, Olah ME, Sun X, Czerwinska E, et al. (2007) TRPM7 channels in hippocampal neurons detect levels of extracellular divalent cations. Proc Natl Acad Sci U S A 104: 16323–16328.
  11. 11. Lu B, Zhang Q, Wang H, Wang Y, Nakayama M, et al. (2010) Extracellular calcium controls background current and neuronal excitability via an UNC79-UNC80-NALCN cation channel complex. Neuron 68: 488–499.
  12. 12. Schneggenburger R, Neher E (2005) Presynaptic calcium and control of vesicle fusion. Curr Opin Neurobiol 15: 266–274.
  13. 13. Huang LY, Neher E (1996) Ca(2+)-dependent exocytosis in the somata of dorsal root ganglion neurons. Neuron 17: 135–145.
  14. 14. Zhang C, Xiong W, Zheng H, Wang L, Lu B, et al. (2004) Calcium- and dynamin-independent endocytosis in dorsal root ganglion neurons. Neuron 42: 225–236.
  15. 15. Chen XW, Feng YQ, Hao CJ, Guo XL, He X, et al. (2008) DTNBP1, a schizophrenia susceptibility gene, affects kinetics of transmitter release. J Cell Biol 181: 791–801.
  16. 16. Zhang C, Zhou Z (2002) Ca(2+)-independent but voltage-dependent secretion in mammalian dorsal root ganglion neurons. Nat Neurosci 5: 425–430.
  17. 17. Wu B, Wang YM, Xiong W, Zheng LH, Fu CL, et al. (2005) Optimization of a multi-channel puffer system for rapid delivery of solutions during patch-clamp experiments. Front Biosci 10: 761–767.
  18. 18. Gillis KD (1995) in Single Channel Recording,. In: Sakmann B, Neher E, editors. pp. 155–198. (Plenum, New York).
  19. 19. Zhou Z, Misler S (1996) Amperometric detection of quantal secretion from patch-clamped rat pancreatic beta-cells. J Biol Chem 271: 270–277.
  20. 20. Zhou Z, Misler S (1995) Action potential-induced quantal secretion of catecholamines from rat adrenal chromaffin cells. J Biol Chem 270: 3498–3505.
  21. 21. Huang HP, Wang SR, Yao W, Zhang C, Zhou Y, et al. (2007) Long latency of evoked quantal transmitter release from somata of locus coeruleus neurons in rat pontine slices. Proc Natl Acad Sci U S A 104: 1401–1406.
  22. 22. Ellis-Davies GC, Kaplan JH (1994) Nitrophenyl-EGTA, a photolabile chelator that selectively binds Ca2+ with high affinity and releases it rapidly upon photolysis. Proc Natl Acad Sci U S A 91: 187–191.
  23. 23. Xu T, Binz T, Niemann H, Neher E (1998) Multiple kinetic components of exocytosis distinguished by neurotoxin sensitivity. Nat Neurosci 1: 192–200.
  24. 24. Xu T, Rammner B, Margittai M, Artalejo AR, Neher E, et al. (1999) Inhibition of SNARE complex assembly differentially affects kinetic components of exocytosis. Cell 99: 713–722.
  25. 25. Ouyang K, Zheng H, Qin X, Zhang C, Yang D, et al. (2005) Ca2+ sparks and secretion in dorsal root ganglion neurons. Proc Natl Acad Sci U S A 102: 12259–12264.
  26. 26. Ruat M, Molliver ME, Snowman AM, Snyder SH (1995) Calcium sensing receptor: molecular cloning in rat and localization to nerve terminals. Proc Natl Acad Sci U S A 92: 3161–3165.
  27. 27. Han W, Rhee JS, Maximov A, Lin W, Hammer RE, et al. (2005) C-terminal ECFP fusion impairs synaptotagmin 1 function: crowding out synaptotagmin 1. J Biol Chem 280: 5089–5100.
  28. 28. Heyeraas KJ, Haug SR, Bukoski RD, Awumey EM (2008) Identification of a Ca2+-sensing receptor in rat trigeminal ganglia, sensory axons, and tooth dental pulp. Calcif Tissue Int 82: 57–65.
  29. 29. Zhang X, Chen Y, Wang C, Huang LY (2007) Neuronal somatic ATP release triggers neuron-satellite glial cell communication in dorsal root ganglia. Proc Natl Acad Sci U S A 104: 9864–9869.
  30. 30. Zheng H, Fan J, Xiong W, Zhang C, Wang XB, et al. (2009) Action potential modulates Ca2+-dependent and Ca2+-independent secretion in a sensory neuron. Biophys J 96: 2449–2456.
  31. 31. Chen XK, Wang LC, Zhou Y, Cai Q, Prakriya M, et al. (2005) Activation of GPCRs modulates quantal size in chromaffin cells through G(betagamma) and PKC. Nat Neurosci 8: 1160–1168.
  32. 32. Neher E (1998) Vesicle pools and Ca2+ microdomains: new tools for understanding their roles in neurotransmitter release. Neuron 20: 389–399.
  33. 33. Nagy G, Reim K, Matti U, Brose N, Binz T, et al. (2004) Regulation of releasable vesicle pool sizes by protein kinase A-dependent phosphorylation of SNAP-25. Neuron 41: 417–429.
  34. 34. Gad H, Low P, Zotova E, Brodin L, Shupliakov O (1998) Dissociation between Ca2+-triggered synaptic vesicle exocytosis and clathrin-mediated endocytosis at a central synapse. Neuron 21: 607–616.
  35. 35. Artalejo CR, Henley JR, McNiven MA, Palfrey HC (1995) Rapid endocytosis coupled to exocytosis in adrenal chromaffin cells involves Ca2+, GTP, and dynamin but not clathrin. Proc Natl Acad Sci U S A 92: 8328–8332.
  36. 36. Fernandez JM, Neher E, Gomperts BD (1984) Capacitance measurements reveal stepwise fusion events in degranulating mast cells. Nature 312: 453–455.
  37. 37. Burgoyne RD, Handel SE (1994) Activation of exocytosis by GTP analogues in adrenal chromaffin cells revealed by patch-clamp capacitance measurement. FEBS Lett 344: 139–142.
  38. 38. Chow RH, von Ruden L, Neher E (1992) Delay in vesicle fusion revealed by electrochemical monitoring of single secretory events in adrenal chromaffin cells. Nature 356: 60–63.
  39. 39. Wightman RM, Jankowski JA, Kennedy RT, Kawagoe KT, Schroeder TJ, et al. (1991) Temporally resolved catecholamine spikes correspond to single vesicle release from individual chromaffin cells. Proc Natl Acad Sci U S A 88: 10754–10758.
  40. 40. Schneggenburger R, Neher E (2000) Intracellular calcium dependence of transmitter release rates at a fast central synapse. Nature 406: 889–893.
  41. 41. Voets T, Moser T, Lund PE, Chow RH, Geppert M, et al. (2001) Intracellular calcium dependence of large dense-core vesicle exocytosis in the absence of synaptotagmin I. Proc Natl Acad Sci U S A 98: 11680–11685.
  42. 42. Cohen JE, Fields RD (2004) Extracellular calcium depletion in synaptic transmission. Neuroscientist 10: 12–17.
  43. 43. Hofer AM, Brown EM (2003) Extracellular calcium sensing and signalling. Nat Rev Mol Cell Biol 4: 530–538.
  44. 44. Ferry S, Traiffort E, Stinnakre J, Ruat M (2000) Developmental and adult expression of rat calcium-sensing receptor transcripts in neurons and oligodendrocytes. Eur J Neurosci 12: 872–884.
  45. 45. Awumey EM, Howlett AC, Putney JW Jr, Diz DI, Bukoski RD (2007) Ca(2+) mobilization through dorsal root ganglion Ca(2+)-sensing receptor stably expressed in HEK293 cells. Am J Physiol Cell Physiol 292: C1895–1905.
  46. 46. Nemeth EF, Fox J (1999) Calcimimetic Compounds: a Direct Approach to Controlling Plasma Levels of Parathyroid Hormone in Hyperparathyroidism. Trends Endocrinol Metab 10: 66–71.
  47. 47. Xiong Z, Lu W, MacDonald JF (1997) Extracellular calcium sensed by a novel cation channel in hippocampal neurons. Proc Natl Acad Sci U S A 94: 7012–7017.
  48. 48. Petrel C, Kessler A, Dauban P, Dodd RH, Rognan D, et al. (2004) Positive and negative allosteric modulators of the Ca2+-sensing receptor interact within overlapping but not identical binding sites in the transmembrane domain. J Biol Chem 279: 18990–18997.
  49. 49. Faure H, Gorojankina T, Rice N, Dauban P, Dodd RH, et al. (2009) Molecular determinants of non-competitive antagonist binding to the mouse GPRC6A receptor. Cell Calcium 46: 323–332.
  50. 50. Phillips CG, Harnett MT, Chen W, Smith SM (2008) Calcium-sensing receptor activation depresses synaptic transmission. J Neurosci 28: 12062–12070.
  51. 51. Sudhof TC (2004) The synaptic vesicle cycle. Annu Rev Neurosci 27: 509–547.
  52. 52. Tang J, Maximov A, Shin OH, Dai H, Rizo J, et al. (2006) A complexin/synaptotagmin 1 switch controls fast synaptic vesicle exocytosis. Cell 126: 1175–1187.
  53. 53. Moser T, Neher E (1997) Rapid exocytosis in single chromaffin cells recorded from mouse adrenal slices. J Neurosci 17: 2314–2323.
  54. 54. Zhou Z, Misler S, Chow RH (1996) Rapid fluctuations in transmitter release from single vesicles in bovine adrenal chromaffin cells. Biophys J 70: 1543–1552.
  55. 55. Ales E, Tabares L, Poyato JM, Valero V, Lindau M, et al. (1999) High calcium concentrations shift the mode of exocytosis to the kiss-and-run mechanism. Nat Cell Biol 1: 40–44.
  56. 56. Kishimoto T, Kimura R, Liu TT, Nemoto T, Takahashi N, et al. (2006) Vacuolar sequential exocytosis of large dense-core vesicles in adrenal medulla. EMBO J 25: 673–682.
  57. 57. Stanley EF (2000) Presynaptic calcium channels and the depletion of synaptic cleft calcium ions. J Neurophysiol 83: 477–482.
  58. 58. Rusakov DA, Fine A (2003) Extracellular Ca2+ depletion contributes to fast activity-dependent modulation of synaptic transmission in the brain. Neuron 37: 287–297.
  59. 59. Katz B, Miledi R (1967) The timing of calcium action during neuromuscular transmission. J Physiol 189: 535–544.
  60. 60. Neher E, Sakaba T (2008) Multiple roles of calcium ions in the regulation of neurotransmitter release. Neuron 59: 861–872.
  61. 61. Brown EM, Gamba G, Riccardi D, Lombardi M, Butters R, et al. (1993) Cloning and characterization of an extracellular Ca(2+)-sensing receptor from bovine parathyroid. Nature 366: 575–580.
  62. 62. Atlas D (2010) Signaling role of the voltage-gated calcium channel as the molecular on/off-switch of secretion. Cell Signal 22: 1597–1603.
  63. 63. Lerner I, Trus M, Cohen R, Yizhar O, Nussinovitch I, et al. (2006) Ion interaction at the pore of Lc-type Ca2+ channel is sufficient to mediate depolarization-induced exocytosis. J Neurochem 97: 116–127.
  64. 64. Campbell DT, Hille B (1976) Kinetic and pharmacological properties of the sodium channel of frog skeletal muscle. J Gen Physiol 67: 309–323.
  65. 65. Yang DY, Lee JB, Lin MC, Huang YL, Liu HW, et al. (2004) The determination of brain magnesium and zinc levels by a dual-probe microdialysis and graphite furnace atomic absorption spectrometry. J Am Coll Nutr 23: 552S–555S.